heat and mass transfer analogy under turbulent conditions of frying

Transcription

heat and mass transfer analogy under turbulent conditions of frying
HEAT AND MASS TRANSFER ANALOGY UNDER TURBULENT
CONDITIONS OF FRYING
A Thesis Submitted to the
College of Graduate Studies and Research
In Partial Fulfillment of the Requirements
For the Degree of Master of Science
In the Department of Agricultural and Bioresource Engineering
University of Saskatchewan
Saskatoon, Saskatchewan
By
Adefemi Farinu
© Copyright Adefemi Farinu, October 2006. All rights reserved.
PERMISSION TO USE
In presenting this thesis in partial fulfillment of the requirements for a
Postgraduate degree from the University of Saskatchewan, I agree that the
Libraries of this University may make it freely available for inspection. I further
agree to permission for copying of this thesis in any manner, in whole or part, for
scholarly purposes as may be granted by the professor who supervised my
thesis work, or in his absence, by the Head of the Department or the Dean of the
College in which my thesis work was done. It is understood that any copying or
publication or use of this thesis or parts thereof for financial gain shall not be
allowed without my written permission. It is also understood that due recognition
shall be given to me and to the University of Saskatchewan in any scholarly use
which may be made of any material in my thesis.
Requests for permission to copy or to make other use of material in this thesis in
whole or in part should be addressed to:
Head, Department of Agricultural and Bioresource Engineering
University of Saskatchewan
Saskatoon, Saskatchewan
S7N 5A9
CANADA
ii
ABSTRACT
Sweetpotato (Ipomoea batatas) is a popular vegetable across the world.
It is a staple food item of many countries in South America, Africa and Asia
where the population depends on the crop as an important source of energy and
essential nutrients like vitamins A and C, calcium, iron and copper. It is also a
very popular crop in North America. Deep fat frying is one of the favourite
processing methods for sweetpotato. The method is fast and the finished
product is desired for its unique flavour and taste.
The main objective of this study was to establish analogy between
convective heat and mass transfer during frying. The accurate estimation of the
coefficients for both phenomena is challenging. During frying, the rate of heat
transfer from the oil to the food surface is largely controlled by the convective
heat transfer coefficient. This heat transfer coefficient is dependent on the
interaction between the temperature gradient and the drying rate in a frying
process. The temperature gradient and the drying rate in turn partly depend on
the thermophysical properties of the product. In this study, thermophysical
properties of sweetpotato were studied and modeled as a function of moisture
content and temperature. The properties of interest are specific heat capacity,
thermal conductivity, thermal diffusivity and density. A designed deep fat frying
experiment of sweetpotato was carried out under four different oil temperatures
(150, 160, 170 and 180°C) and using three different sample sizes (defined as
ratio of diameter to thickness, D/L: 2.5, 3.5 and 4.0). Convective heat transfer
coefficients under these frying conditions were estimated and computer
iii
simulation based on finite element modeling technique was used to determine
convective mass transfer coefficients. Correlation between heat transfer
coefficient and mass transfer coefficient were investigated with reliable statistical
tool. Effects of sample size, oil temperature and frying time on heat and mass
transfer were also studied.
Specific heat, thermal conductivity and thermal diffusivity of sweetpotato
were all found to increase with increase in temperature and moisture content.
Density decreased with increase in moisture content. Maximum heat transfer
coefficient reached during sweetpotato frying was in the range of 700-850
W/m2°C. Heat transfer coefficient of sample during frying increased with
increase in frying oil temperature but decreased with increase in sample size.
Same trend for heat transfer coefficient was observed for effects of oil
temperature and sample size on mass transfer coefficient. Maximum mass
transfer coefficient reached during sweetpotato frying was in the range of 4×10-6
to 7.2×10-6 kg/m2.s. No general relationship was established between heat
transfer coefficient and mass transfer coefficient during frying but a relationship
was established between maximum heat transfer coefficient and maximum
mass transfer coefficient. A trend was also observed between maximum heat
transfer coefficient and the corresponding mass transfer coefficient at that point.
iv
ACKNOWLEDGEMENT
I am indebted to my supervisor Dr. Oon-Doo Baik. I owe to him the
opportunity for this study in the first place. His extreme patience, constant
encouragement and infectious enthusiasm made all the difference in completing
this study.
I would like to express my gratitude to my graduate committee members
Dr. Venkatesh Meda and Dr. Lope Tabil Jr. for sharing their wealth of knowledge
with me and making themselves available for guidance at various stages of this
program.
I would also like to thank Natural Sciences and Engineering Research
Council of Canada (NSERC) for financial support and Agriculture and Agri-Food
Canada for providing equipment for Specific heat measurement. I would like to
thank the department of Agricultural and Bioresource Engineering, University of
Saskatchewan as a whole and my colleagues for the opportunity, academic and
moral support. I must mention Wayne Morley, Louis Roth, Toni Schleicher and
Bill Crerar for their technical helps in my experiments, and Pat Hunchak for
christening me after a Legend (Michael Jordan) on my first day in the
department.
My appreciation goes to friends in Canada: Yinka, Samuel, Jide, Taiwo,
Lekan, Tope, Paul, Ebenezer, Dapo and friends in Nigeria: Noble brothers.
Specific mention of Gbola Oladokun is necessary.
My utmost thanks go to Jehovah. He is indeed faithful.
v
DEDICATION
This thesis work is dedicated to my family: Ajibade, Bolanle, Adeyinka,
Adedoyin, Adedunmola, Oluwaseun and also to Juliet.
vi
TABLE OF CONTENTS
PERMISSION TO USE…………………………………………………………….. ii
ABSTRACT…………………………………………………………………………. iii
ACKNOWLEDGEMENTS………………………………………………………..... v
DEDICATION…………………………………………………………………….…. vi
TABLE OF CONTENTS…………………………………………………………... vii
LIST OF FIGURES……………………………………………………………….... x
LIST OF TABLES…………………………………………………………………... xii
NOMENCLATURE…………………………………………………………………. xiii
CHAPTER 1 INTRODUCTION……………………………………………………..1
1.1 Background……………………………………………………………... 1
1.2 Statement of problem………………………………………………….. 3
1.3 Objectives……………………………………………………………….. 3
CHAPTER 2 LITERATURE REVIEW…………………………………………......5
2.1 Sweetpotato…………………………………………………………….. 5
2.2 Thermophysical properties……………………………………………. 6
2.2.1 Specific heat capacity……………………………………………….. 7
2.2.2 Thermal conductivity………………………………………………..
10
2.2.3 Thermal Diffusivity………………………………………………….. 13
2.2.4 Density……………………………………………………………….
15
2.2.5 Moisture Diffusivity………………………………………………….
17
2.3 Deep fat frying…………………………………………………………
18
2.3.1 The frying process…………………………………………………..
19
2.3.2 Heat transfer………………………………………………………...
20
2.3.3 Moisture transfer…………………………………………………….
22
2.4 Mathematical modeling……………………………………………….
25
2.5 Heat and mass transfer analogy……………………………………
33
vii
2.6 Summary… … … … … … … … … … … … … … … … … … … … … … … …
38
CHAPTER 3 MATERIALS AND METHODS… … … … … … … … … … … … … … 39
3.1 Materials… … … … … … … … … … … … … … … … … … … … … … … … ... 39
3.2 Density measurement… … … … … … … … … … … … … … … … … … … 39
3.2.1 Multipycnometer… … … … … … … … … … … … … … … … … … … … .. 40
3.2.2 Mechanistic model… … … … … … … … … … … … … … … … … … … ... 42
3.3 Specific heat measurement… … … … … … … … … … … … … … … … .. 45
3.3.1 Mechanistic model… … … … … … … … … … … … … … … … … … … .
46
3.3.2 DSC… … … … … … … … … … … … … … … … … … … … … … … … … .. 46
3.4 Thermal conductivity… … … … … … … … … … … … … … … … … … … . 49
3.5 Thermal diffusivity… … … … … … … … … … … … … … … … … … … … . 51
3.6 Moisture diffusivity… … … … … … … … … … … … … … … … … … … … . 51
3.7 Deep fat frying… … … … … … … … … … … … … … … … … … … … … … 52
3.7.1 Sample preparation… … … … … … … … … … … … … … … … … … … 52
3.7.2 Sample holder… … … … … … … … … … … … … … … … … … … … … . 53
3.7.3 Experimental Procedure… … … … … … … … … … … … … … … … … 54
3.7.4 Temperature measurement… … … … … … … … … … … … … … … .. 57
3.7.5 Moisture content measurement… … … … … … … … … … … … … … 57
3.8 Heat transfer coefficient… … … … … … … … … … … … … … … … … … 58
3.9 Computer simulation of heat and mass transfer… … … … … … … ..
59
3.9.1 COMSOL™ software… … … … … … … … … … … … … … … … … …
59
3.9.2 Background, assumptions and approach to simulation … .… …
61
3.9.3 Criterion for best fit… … … … … … … … … … … … … … … … … … … … 63
3.9.4 Governing equations, initial conditions and boundary… … … … . 63
conditions
CHAPTER 4 RESULTS AND DISCUSSIONS… … … … … … … … … … … … … 67
4.1 Density… … … … … … … … … … … … … … … … … … … … … … … … … 67
4.1.1 Measured density.… … … … … … … … … … … … … … … … … … …
67
4.1.2 Mechanistic model… … … … … … … … … … … … … … … … … … …
68
4.2 Specific heat capacity… … … … … … … … … … … … … … … … … … ..
70
viii
4.2.1 DSC… … … … … … … … … … … … … … … … … … … … … … … … … .
70
4.2.2 Mechanistic model… … … … … … … … … … … … … … … … … … …
72
4.3 Thermal conductivity… … … … … … … … … … … … … … … … … … …
74
4.4 Thermal diffusivity… … … … … … … … … … … … … … … … … … … …
76
4.5 Moisture content during frying… … … … … … … … … … … … … … …
78
4.6 Temperature during frying… … … … … … … … … … … … … … … … … 81
4.7 Heat transfer coefficient… … … … … … … … … … … … … … … … … ..
83
4.8 Computer simulation results… … … … … … … … … … … … … … … … . 85
4.8.1 Comparison of temperature profiles of experiment and
Simulation… … … … … … … … … … … … … … … … … … … … … ........ 87
4.8.2 Comparison of moisture content profiles of experiment and
Simulation… … … … … … … … … … … … … … … … … … … … … … .... 89
4.8.3 Mass transfer coefficient result… … … … … … … … … … ..............
92
4.9 Correlation between heat transfer coefficient, h and mass
transfer coefficient, hm… … … … … … … … … … … … … … … … … … .. 93
4.10 Effects of experimental factors on h and hm… … … … … … … … … 94
CHAPTER 5 SUMMARY, CONCLUSIONS AND RECOMMENDATIONS… .. 96
5.1 Summary and conclusions… … … … … … … … … … … … … … … … … 96
5.2 Recommendations… … … … … … … … … … … … … … … … … … … … .. 97
REFERENCES… … … … … … … … … … … … … … … … … … … … … … … … … … 99
APPENDICES… … … … … … … … … … … … … … … … … … … … … … … … … … … .104
ix
LIST OF FIGURES
Figure 3.1 Multipycnometer used to measure density of sweetpotato… … … .
41
Figure 3.2 DSC 2010 used to measure specific heat of sweetpotato … … …
48
Figure 3.3 Sweetpotato sample sizes and cylindrical borers used to cut
samples… … … … … … … … … … … … ..… … … … … .… … … … ...........
53
Figure 3.4 Sample holder used for frying with thermocouples attached… … … 54
Figure 3.5 Deep fryer used for sweetpotato frying… … … … … … … … … … … .... 56
Figure 3.6 Sweetpotato sample in the oil during frying… … … … … … … … … … 57
Figure 3.7 Crust and core samples of sweetpotato after frying… … … … … …
58
Figure 3.8 Mesh generation in the sweetpotato sample domain… … … … …
61
Figure 3.9 Sweetpotato sample disc geometry… … … … … … … … ..… … … …
62
Figure 4.1 Density of sweetpotato as a function of moisture content determined
from multipycnometer… … … … … … … … … … … … … … … … … … ..
67
Figure 4.2 Density of sweetpotato as a function of moisture content
and temperature determined from mechanistic model… … … … .
69
Figure 4.3 Specific heat of sweetpotato as a function of temperature
and moisture content determined from DSC.… … … … … … … …
72
Figure 4.4 Specific heat of sweetpotato as a function of temperature
and moisture content from mechanistic model… … … … … … … .
74
Figure 4.5 Thermal conductivity of sweetpotato as a function of moisture
content and temperature determined from mechanistic model… .. 76
Figure 4.6 Thermal diffusivity of sweetpotato as a function of moisture content
and temperature determined from mechanistic model… … … … … .. 78
x
Figure 4.7 Moisture content variation among sample sizes (whole sample)
fried at 180°C… … … … … … … … … … … … … … … … … … … … … … ... 80
Figure 4.8 Moisture content of big sample (D/L=4) during frying at four
different oil temperatures… … … … … … … … … … … … … … … … … ... 80
Figure 4.9 Temperature of medium sample (D/L=3.5) during frying at 150°C... 82
Figure 4.10 Temperature of medium sample (D/L=3.5) during frying at 180°C. 83
Figure 4.11 Heat transfer coefficient of medium sample (D/L=3.5) fried at
various temperatures during sweetpotato frying… … … … … … … … 84
Figure 4.12 Heat transfer coefficient among various sample sizes fried at 150°C
during sweetpotato frying… .… … … … … … … … … … … … … … … … . 85
Figure 4.13 Simulated temperature profile of medium sample (D/L=3.5) fried at
170°C… … … … … … … … … … … … … … … … … … … … … … … … … .. 86
Figure 4.14 Simulated moisture content of medium sample (D/L=3.5) fried at
170°C… … … … … … … … … … … … … … … … … … … … … … … … … .. 86
Figure 4.15 Comparison of surface temperature profiles of experiment
and simulation for medium sample fried at 160°C… … … … … … … 88
Figure 4.16 Comparison of centre temperature profiles of experiment
and simulation for medium sample fried at 160°C… … … … … … ... 89
Figure 4.17 Comparison of sample surface moisture content of experiment and
simulation for medium sample fried at 150°C… … … … … … .… … . 91
Figure 4.18 Comparison of whole sample moisture content profiles of
experiment and simulation for medium sample fried at 150°C .… 91
Figure 4.19 Mass transfer coefficient determined during sample frying… ....… 92
xi
LIST OF TABLES
Table 3.1 Models used in this study for major components of foods… ..........
44
Table 3.2 Percentage composition of major food components in sweetpotato..45
Table A1.Results of sweetpotato density determined by multipycnometer… 104
Table A2. Results of sweetpotato specific heat determined by DSC… … .… ...104
Table A3. Measured moisture content of sweetpotato during frying… … … … 105
Table A4. Measured moisture content of sweetpotato during frying… … … … ..106
Table A5. Measured temperature of sweetpotato sample during frying… … .. 107
Table A6. Measured temperature of sweetpotato sample during frying… … .. 108
Table A7. Average heat transfer coefficient during sample frying… … … … … ..109
Table A8. Average heat transfer coefficient during sample frying… … … … … ..109
Table A9. Determined average mass transfer coefficient during frying........... 110
Table A10. Determined average mass transfer coefficient during frying...… .. 111
Table B1. Summarised ANOVA table for h and hm correlation at maximum h.112
Table B2. Summarised ANOVA table for maximum h and hm correlation… ...112
Table B3. Summarised ANOVA table for h correlation with factors of frying… 112
Table B4. Summarised ANOVA table for hm correlation with factors of frying..112
xii
NOMENCLATURE
A = area (m2)
c = concentration (kg/m3)
Cm = specific moisture capacity (kgmoisture/kgsample)
Cp = specific heat (J/kg.K)
d = sample diameter to thickness ratio i.e. D/L
D = mass diffusivity (m2/s)
h = heat transfer coefficient (W/m2.K)
hm = mass transfer coefficient (kg/m2.s)
.
I = volumetric expansion (kg/m3.s)
k = thermal conductivity (W/m.K)
ki = the thermal conductivity of ith phase
km = moisture conductivity (kg/m.s)
L = length of material (m)
m = moisture content (wet basis)
mw = moisture content (g water/g solid).
M = mass of material (kg)
Mf = oil content (% wet basis)
mfep = equilibrium fat content (mass fraction, dry basis)
n = mass flux (kg/m2)
KL = fat conductivity (m/s)
P = Pressure (Pa)
Pr = Prandtl number
xiii
q = heat flux (W/m2)
Q = heat lost or gained (J)
R = radius (m)
r = radial distance (m)
Sc = Schmidt number
S = saturation
T = temperature (°C)
t = time (s)
To = oil temperature (°C)
Ts = surface temperature (°C)
U = velocity (m/s2)
V = product volume (m3)
v = kinematic viscosity (m2/s)
Vp = volume of the sample (cm3)
Vc = volume of the cell containing the sample (cm3)
VR = volume of the reference cell (cm3)
x = material thickness (m)
Xiw = mass fraction of food component i.
Xiv = volume fraction of component i
1 = thermal diffusivity (m2/s)
2 = dynamic viscosity (kg/m.s)
3
4
= dimensionless distance
5 = density (kg/m3)
xiv
φ = porosity
6abs = absolute temperature (K)
7 = latent heat of evaporation of water (J/kg)
ε i = Volume fraction of specie i ( mi3 / mt3 )
Subscripts: v, w, a, o, g, s, p, and eff are vapour, water, air, oil, gas (vapour+air),
solid matrix (surface), partial pressure, and effective, respectively.
xv
1. INTRODUCTION
1.1 Background
Sweetpotato (Ipomoea batatas) is a crop whose large starchy sweet
tasting tuberous root is an important vegetable. It is not particularly related to the
regular potato, which is sometimes called the Irish potato to distinguish it. The
two crops are from different families. In rating 58 vegetables by adding up the
percentages of recommended daily allowance (RDA) for six nutrients (Vitamin A
and C, folate, iron, copper and calcium), plus fibre, the Nutrition Action Health
Letter rated sweetpotato the highest with 582 points; its nearest competitor, a
raw carrot came in at 434. Also, the Centre for Science in the Public Interest
rated the relative nutritional value of common vegetables and once again, the
sweetpotato came out on top with a score of 184, as compared with a similarly
prepared regular potato, which scored only 83 points (Rodriguez et. al., 1975).
Sweetpotato’s processing has not been fully studied yet despite its popularity
and importance all over the world.
Deep fat frying is a very important and popular method of food processing.
The snack industry in North America and all other part of the world is a multibillion dollar venture and deep fat fried foods are at the center of this industry.
Despite health concerns for fried foods in terms of high cholesterol content of
1
some oil used in deep frying, they are still craved for the unique flavour and
speed of the process.
Deep fat fried sweetpotato is popular all over the world. In this process,
sweetpotato slices are cooked by immersion in an edible fat or oil at temperature
between 150 and 200°C. During the frying process, heat and mass transfer
within the product occurs simultaneously. Heat and oil are transferred from the
frying medium to the sweet potato while moisture is transferred from the product
to the frying oil. The quality of the fried product depends on the frying process.
Frying is a very turbulent process (Re > 5 x 105). Turbulence is
associated with the existence of random fluctuations in the fluid. Transport of
heat, mass and momentum in a turbulent boundary layer is attributed to motion
of eddies, which are small portion of fluid in the boundary layer that move about
for a short time before losing their identity. Because of this motion, the transport
of energy mass and momentum is greatly enhanced. The existence of
turbulence therefore is advantageous in providing increased heat and mass
transfer rates. However, it makes the process more complicated to describe
theoretically. In modeling the frying process, there is a need to determine the
convective boundary conditions in heat and mass transfer. The factors that
govern the rate of heat and mass transfer during frying are the heat transfer
coefficient and mass transfer coefficient and they are interrelated for a particular
process. The oil temperature during frying as well as the thermal and physical
characteristics of the product and those of the oil is important in determining
these coefficients. Most deep fat frying studies in literature focussed on
determining heat transfer coefficient. Studies that actually quantified the
2
corresponding mass transfer coefficient and investigated the link between the
two coefficients are very scarce.
1.2 Statement of the problem
The relationship between heat transfer coefficient and mass transfer
coefficient is very important in determining the rate of heat and mass transfer
during deep fat frying. Since mass transfer in a food product is driven by the
heat transferred to the product, both are interwoven and establishing the
relationship between them is vital to understanding the dynamics of the process
and predicting the rate. Studies on heat and mass transfer analogy under both
laminar and turbulent conditions exist in literature. However, no model has been
developed yet for the special case of frying. The turbulent nature of the process
by the bubbling action and the structural changes that accompanies a fried
product make the process both fascinating and challenging. Available
information in literature on processing of sweetpotato does not justify the
importance of the crop. This thesis work hopes to fill this gap.
1.3 Objectives
The general objective of this study is to develop heat and mass transfer
analogy during deep fat frying of sweetpotato based on a finite element method
computer simulation.
The specific objectives are:
1. to model the thermophysical properties of sweetpotato as a function of
product temperature and moisture content;
3
2. to determine the convective heat transfer coefficient of deep fat fried
sweetpotato from temperature and moisture content profiles of a
designed experiment;
3. to simulate sweetpotato frying and determine convective mass transfer
coefficient using a finite element method based numerical model; and
4. to develop an empirical correlation between heat transfer coefficient and
mass transfer coefficient during frying based on geometries and oil
temperatures applied and investigate the effects of these factors on the
coefficients.
4
2. LITERATURE REVIEW
This chapter reviews research work along the areas of this study found in
literature. It is basically a review on sweetpotato, thermophysical properties
measurement, deep fat frying process, mathematical modeling and simulation of
deep fat frying process and heat and mass transfer analogy.
2.1 Sweetpotato
Sweetpotato (Ipomoea batatas) belongs to the Convolvulaceae or
morning-glory family and is a dicot (having a two-leafed embryo). The genus
consists of more than 400 species of which, more than 200 occur in tropical
America (Charney and Seelig, 1967). It is a tuberous-rooted perennial crop. Its
stems are usually prostrate and slender. The root does not have eyes or scars
as found on some other roots and tubers, but it possesses the ability to develop
adventitious buds on sprouts or vine cuttings, which is advantageous in
reproducing the crop by vegetative means (Charney and Seelig, 1967). History
has it that Columbus and his shipmates found the native-American plant and
mentioned it on their fourth voyage (1502-1504). The Incas of South America
and Mayas of Central America grew several varieties and called the plant cassiri.
Sweetpotato reached St. Thomas off the African coast before 1574. It is
5
believed that early Spanish explorers took the sweetpotato to the Philippines
and the East Indies, after which Portuguese voyagers carried it to India, China
and Malaya (Charney and Seelig, 1967).
Both yellow and white types of sweetpotato exist, the colour being of the
flesh. The yellow type is preferred because of the attractive colour, good flavour
and cooking qualities (Rodriguez et. al. 1975). However it is not as sweet as the
white type. Sweetpotato is an exceptionally rich source of Vitamin A (7100
IU/100 g). It also has appreciable quantities of ascorbic acid, thiamine, riboflavin,
niacin, phosphorus, iron and calcium (U.S.D.A, 1984; Picha, 1985).
Sweetpotatoes are usually consumed after baking, boiling, steaming, frying or it
may be candied with syrup, sliced into chips, or pureed. Deep fat frying is one of
the popular methods of processing sweetpotato. It is used to produce French
fries or chips.
2.2 Thermophysical properties
The knowledge of thermal and physical properties is essential for
thermodynamic research and modeling the heat treatment of foods including
vegetables. This is because the properties, to a large extent, determine the rate
of the heat transfer process. Previously, constant average values of thermal
properties were used in analyses of food processes, which have lead to
inaccuracies since these properties actually change during the process.
However, modern analytical techniques have made it possible to accommodate
these dynamic changes. The most important thermo-physical properties in food
6
processing, specific heat capacity, thermal conductivity, density and thermal
diffusivity of food materials depend mostly on the food’s composition,
temperature and density (Choi and Okos, 1985). A lot of work has been done on
thermal properties of vegetables; Irish-potato for instance, but very little work
has been done on thermal properties of sweetpotato. Related research works
found in literature are discussed in this section
2.2.1 Specific heat capacity
The specific heat of a substance represents the variation of temperature
with the amount of heat stored within the substance. It depends on the nature of
the heat addition process in terms of either a constant pressure or a constant
volume process. However since pressure change in heat transfer problems of
food materials are usually very small, the specific heat at constant pressure is
most often considered. Specific heat is the ratio of heat loss or gained to
temperature change for a unit mass. Mathematically, it can be represented as;
Cp =
Q
M∆T
(2.1)
Where
Q = heat lost or gained (J),
M = mass of product (kg),
8T = temperature change in the food material (K),
Cp = specific heat (J/kg.K).
7
The need for reliable specific heat data for food materials has been
recognised for a long time. Siebel (1892) proposed that the specific heat of food
materials such as fruits, eggs, vegetables and meat could be assumed equal to
the specific heat of water and that of the solid matter in combination with water.
Siebel (1892) proposed the following equation for food at temperature above
freezing point:
Cp = 0.008m + 0.2
(2.2)
Where:
Cp = specific heat capacity of the food (Btu/lb.°F)
m = moisture content of the material (percentage wet basis);
0.2 = a constant assumed to be specific heat of dry solid in the food.
Stitt and Kennedy (1945) suggested that the constant in the Siebel
equation (equation 2.2) be changed from 0.2 to 0.32 for temperature range of
32°F to 77°F and 0.45 for temperature range of 77°F to 149°F. Sweat (1986),
however, noted that both Siebel’s linear model and the modification made by
Stitt and Kennedy (1945) gave significant deviation from experimental results at
lower temperature. Hence the application of the models was divided into two
ranges, one above M.C of 50% (w.b.) and one below 50% (w.b.).
Choi and Okos (1983) developed a specific heat model which takes into account
the individual contribution of major food components:
Cp = 4180Xww +1711 Xpw +1928 Xfw +1547 Xcw + 908 Xaw
Where:
8
(2.3)
Cp = specific heat of food (J/kg.K)
Xw= mass fraction of each component.
Subscripts w, p, f, c and a are water, protein, fat, carbohydrate and ash,
respectively.
Methods of mixtures and differential scanning calorimetry are the most
common experimental methods used for specific heat determination of food
materials. Hwang and Hayakawa (1979) used a calorimeter to measure the
specific heat of food materials. The authors avoided direct contact between food
and the exchange medium in the calorimeter. Two rubber O-rings were
introduced into the calorimeter to prevent water loss by evaporation. Gupta
(1990) reported a simpler calorimeter than that of Hwang and Hayakawa (1979).
In this study, the test capsule was made of copper tube with rubber stoppers at
both end of the tube. Gupta (1990) mentioned that the specific heat above
100°C can be measured by using vegetable oil or mineral oil as the heating
medium in place of water.
The method of differential scanning calorimetry, DSC is more advanced
and has enjoyed a wide application in determining the specific heat of food
materials. Moreira et. al. (1995) used DSC to determine the specific heat of
tortilla chips as a function of frying time. Scanning was conducted at a heating
rate of 12°C/min. Cp of tortilla chip from this study ranged from 2560 to 3360
J/kg.K. Kramkowski and co-worker (2001) determined the specific heat of garlic
as a function of temperature at different moisture content levels. Specific heat
was found to increase with temperature ranging from 2400 to 4100 J/kg.K. Tang
9
and co-workers (1991) used DSC to determine the specific heat capacity of lentil
seeds. Specific heat values of lentil seed increased quadratically with moisture
content over the range of 2.1 to 25.8% w.b. and a linear increase with
temperature varying between 10 and 80°C. Wang and Brennan (1993)
determined specific heat of potato using DSC and model Cp as a function of
moisture content and temperature. Cp of potato was reported to increase
quadratically with moisture content over the range 0-4.13 (g water/g solid) and
increased linearly with temperature varying from 40 to 70°C. Model proposed for
Cp of potato is given in equation (2.4)
Cp = 0.406 +0.00146T + 0.203mw – 0.0249mw2
(2.4)
Where:
Cp = Specific heat (cal/g.°C)
T = Temperature (°C)
mw = moisture content (g water/g solid).
2.2.2 Thermal conductivity
Thermal conductivity, k of a material gives the amount of heat that will be
conducted per unit time through a unit thickness of that material, if a unit
temperature gradient exists across that thickness. Thermal conductivity can be
either determined experimentally or through mathematical estimation. Woodside
and Messmer (1961) and Brailsford and Major (1958) both determined the
maximum and minimum thermal conductivity of a two-phase system as series
(two phases are thermally in series with respect to the direction of heat flow) and
10
parallel (two phases are thermally in parallel with respect to the direction of heat
flow) distribution.
Series model is represented by:
v
n
1
= ∑ Xi
k i =1 k i
(2.5)
While the parallel model is represented by:
n
k = ∑ Xi k i
v
(2.6)
i =1
Where:
Xiv = the volume fraction of ith component phase (m3)
ki = the thermal conductivity of ith phase (W/m.K)
k = thermal conductivity of the product (composite medium) (W/m.K).
Murakami and Okos (1989) considered thermal properties of 15 different
food powders from literature and observed that all data points fall on the range
defined by values predicted by series and parallel models. It was observed that
the parallel model defined the upper limit while the series model defined the
lower limit of the predictions.
Generally, there are two broad methods of experimental measurement of
thermal conductivity in literature. The steady-state methods are based on the
Fourier’s law of steady state heat conduction. The method is mathematically
simple and is quite accurate for dry, granular and solid food. The major
disadvantages of steady-state methods are the long time required to attain
steady-state conditions and the possibility of moisture migration due to
temperature gradient across the material. Transient methods of thermal
11
conductivity measurement make use of the governing equations for the
temperature history in a solid body. Transient methods require a short time for
measurement, which also reduces the chances of moisture migration.
Both steady-state and the transient methods have been used extensively
in food applications. Saravacos and Pilsworth (1965) used the guarded hot plate
for thermal conductivity measurement of several freeze-dried gels and
determined k from the following equation:
k=
qx
A∆T
(2.7)
Where:
k = thermal conductivity (W/m.K)
q = rate of heat input (W)
x = material thickness parallel to heat flow (m)
8T = temperature change (°C)
A = contact area normal to direction of heat flow (m2).
Farral and co-workers (1970) determined the thermal conductivity of
various types of spray-dried and drum-dried powdered milks using the
concentric cylinder method and the following equation:
r 
q ln 2 
 r1 
k=
2πL(T1 − T2 )
(2.8)
Where:
L = length of the cylinder (m),
T1 and T2 = the temperature of the material (K) at radii r1 and r2 (m) respectively,
12
q and k are as defined for equation (2.7).
Kazarian and Hall (1965) used the line source method to measure the
thermal conductivity of yellow dent corn and soft white wheat. k was estimated
using equation (2.9):
k=
t
q
. ln 2
4πL(T1 − T2 ) t1
(2.9)
Where:
T1 and T2 = the temperature (K) at the time t1and t2 (s) after heating respectively,
q = rate of heat input (W),
k = thermal conductivity (W/m.K),
L = length of material (m).
Wang and Brennan (1992) measured thermal conductivity, k of potato
using line-source probe. Thermal conductivity was determined at various
moisture contents (0-4.7 g water/g solid) and in temperature range of 40-70°C. k
of potato was reported to decrease with moisture content and was correlated
with moisture content by a semi-logarithmic equation. The authors reported that
temperature had little effect on k in the temperature range used. Califano and
Calvelo (1991) estimated thermal conductivity of potato from heat penetration
into cylindrical samples between 50 and 100°C. k was found to range from
0.545 to 0.957 W/m°C for potato with specific gravity of 1070 kg/m3 and
moisture content of 80%. k also varied with temperature quadratically.
2.2.3 Thermal Diffusivity
The rate at which heat diffuses through a material is determined by
13
thermal diffusivity123 of the material. The property gives a measure of how
quickly a material’s temperature will change when the material is heated or
cooled. It gives the ratio of conducted energy to stored energy in a material.
Thermal diffusivity can either be determined from direct measurement or
estimated from thermal conductivity, density and specific heat data. For direct
methods, thermal diffusivity is usually determined from the solution of onedimensional unsteady state heat transport equation with the appropriate
boundary conditions for infinite and finite bodies by analytical methods and for
irregular bodies by numerical methods. Indirect estimation of thermal diffusivity
needs considerable time and different instruments since a variety of properties
have to be determined independently. Kazarian and Hall (1965) directly
measured thermal diffusivity of yellow dent corn and soft white wheat. Arce and
co-workers (1981) determined the thermal diffusivity of defatted soy flour
indirectly from thermal conductivity, specific heat and mass density data:
α=
k
ρC p
(2.10)
Where:
k = thermal conductivity (W/m.K),
5 = mass density (kg/m3),
Cp = specific heat capacity (J/kg.K),
1 = thermal diffusivity (m2/s).
Drouzas and co-workers (1991) used direct (probe method) and indirect
method to determine the thermal diffusivity of granular starch and found that
indirect method yielded more accurate values than the direct measurement as
14
concluded from F-test at 5% significance level. Wadsworth and Spadaro (1969)
reported a thermal diffusivity value of 1.1 × 10-7 m2/s for sweetpotato. The
authors heated sweetpotato in a constant temperature water bath at 55, 70, 80
and 90°C and calculated thermal diffusivity from the experimental timetemperature curve. It was further reported that t thermal diffusivity of
sweetpotato increased with temperature while cooking due to increase in
thermal conductivity and changes in starch structure, mainly, starch
gelatinization. Thermal diffusivity, however, started decreasing with temperature
increase after 90°C as starch cells in sweetpotato became distended, weakened
and cell walls started to separate with starch molecules being hydrolyzed.
2.2.4 Density
Density is the ratio of mass to volume of a material. Density of food
products is an important property in analyzing food processing unit operations. It
is closely related to porosity and moisture content of food. True density is the
density of a pure substance or a material calculated from its components’
densities considering conservation of mass and volume. Most density
measurement applications usually borders around volume measurement since
density is the ratio of mass to volume. Measuring the true volume of food
products is challenging due to the irregularity in shapes or characteristic
dimension of most biomaterials and voids in the product that had to be
accounted for. A method that has been used for measuring apparent density of
irregular-shaped product is buoyant force determination (Rahman, 1995). In this
15
procedure, a sample is suspended in a liquid of known density and the weight of
the sample inside the liquid is measured. Buoyant force is the difference
between the mass in air and the mass in liquid. Apparent density 5ap is then
determined from equation (2.11):
ρ ap = ρ l
W
G
(2.11)
Where:
W and G = weight of sample in air and buoyant force respectively (N)
5l = density of liquid (kg/m3)
5ap = apparent density of sample (kg/m3).
For most food products, there is exchange of solid, liquid or gas from the
sample when suspended in a liquid and this could be prevented by enclosing the
sample in cellophane or by coating a thin layer of varnish or wax on the sample
(Rahman, 1995). Volume of a sample can also be determined by measurement
of liquid volume displaced using a specific gravity bottle. Mohsenin (1970)
mentioned the characteristics of a liquid preferred when using this method as
little tendency to soak sample, smooth flow, low solvent action, high boiling point,
stable and low specific gravity when exposed to atmosphere. An example of
such liquid is toluene. Gas pycnometer is a good method of measuring porous
and non-porous solids. The device measures volume by using high pressure
gas. Sabapathy (2005) determined particle density of Kabuli chickpea with a gas
multipycnometer at seven levels of moisture content. The author reported that
density decreased with increased moisture content.
16
2.2.5 Moisture Diffusivity
Moisture diffusivity is a measure of a product’s tendency to produce
entropy when it is disturbed from equilibrium by imposition of a concentration
gradient. It is defined as the proportionality constant between a flux and a driving
flux (Kestin and Wakeham, 1988). For any drying application an example of
which is frying, the internal moisture migrates towards the external surface of the
product by means of a number of mechanisms such as diffusion, capillarity and
sequences of evaporation-condensation. Generally, two or more of these
mechanisms occur simultaneously, making rigorous theoretical modeling of
drying kinetics quite difficult due to variable transfer phenomena coefficients.
Fick’s second law of diffusion (equation 2.12) is widely used to gather all of the
mass transfer mechanisms into one equivalent moisture transport coefficient
(Derdour and Desmorieux, 2004).
∂c ∂  ∂c 
= D 
∂t ∂x  ∂x 
(2.12)
Where:
c = concentration (kg/m3)
t = time (s)
x = material thickness along direction of mass transfer (m)
D = diffusion coefficient (m2/s)
There is no standard method for moisture diffusivity estimation, although
most of the available methods are based on Fick’s laws of diffusion. The
differences among the methods include the kind and the conditions of
experiments used (simple or sophisticated setups, permeation, sorption or
17
drying operation, various specimen geometries) and the treatment of
experimental data. Most of the moisture diffusivity data available originate from
the method of drying curves. What is usually meant by a drying curve is the
representation of change in moisture content of a food specimen with time or the
change in the drying rate with moisture content. Depending on the material and
drying conditions, the drying curve may adopt different shapes. The relationship
between diffusivity and the product temperature is often of the Arrhenius type. D
values for potato exist in literature. For temperature range of 418K to 458K,
moisture diffusivity of potato is given by equation (2.13) (Rice and Gamble,
1989):
 2911 

D = 11.04 exp −
 θ abs 
(2.13)
Where:
D = moisture diffusivity (m2/s)
6abs = absolute temperature (K)
2.3 Deep Fat Frying
Deep fat or immersion frying is defined as the process of cooking foods
by immersing them in edible fat or oil, which is at a temperature above the
boiling point of water, usually 150-200°C (Farkas et al., 1996a). The process
involves simultaneous heat and mass transfer which makes it a complex
process to study and analyse.
2.3.1 The frying process
18
Deep fat frying as a method of food processing combines high processing
speed with good product appearance, although lower yield and higher fat
content was also reported for the method. Deep fat frying is a very fast method
of food processing among conventional heat transfer methods. Bengtsson and
Jacobsson (1974) reported deep fat frying at oil temperatures normally used for
the process as the fastest of three methods: deep fat frying, IR heating and panfrying at normal operating conditions when meat patties of 10 mm thickness was
fried.
Deep fat frying is considered a moving boundary problem, where a
previously non-existent crust region develops on the food surface and increases
in thickness inwards during frying while the core region decreases with frying
time. The process was broken down into four stages by Farkas and co-workers
(1996a). The first stage, known as initial heating, is the period of time within
which the surface of the product is heated from its initial temperature to the
boiling point of water. This phase is usually short and a negligible amount of
water is lost from the food. The second stage, surface boiling, is noted for rapid
loss of surface free moisture, an increase in surface heat transfer coefficient and
beginning of crust formation. The falling rate stage represents the period of time
during which the bulk of the moisture is lost. It is the longest of the stages and
the temperature of the core region approaches that of boiling point of water.
Bubble end point is the final stage of frying. It describes the apparent end of
moisture loss from the product during frying.
19
The following factors were found to govern the process rate in deep fat
frying and also affect the quality of the fried product: the thermophysical
properties of the food and those of the oil, the temperature of the oil, the
geometry of the food and finally, the processing conditions that are likely to lead
to the degradation of the oil in the process (Moreira et. al., 1995a).
2.3.2 Heat Transfer
During deep fat frying of food, heat is transferred from the oil to the
surface of the food immersed in it by convection and onward from the food
surface to its geometric centre by a combination of conduction and advection.
Liquid water moves from inside the food outwards and on reaching an
established evaporation front, turns to vapour and leaves surface as vapour.
Once the moisture on the surface of a food has been removed in a deep fat
frying process, a crust begins to form and its thickness increases over the
duration of frying. The crust reduces the rate at which water is vaporized and
food is cooked (Farkas et al., 1996a). The rate of temperature change within the
core region is only slightly affected by the external oil temperature due to an
isotherm at a slightly elevated water boiling point at the crust/core interface
(Farkas et. al., 1994). Hubbard and Farkas (1999) divided frying process into
two phases with bubbling action: non-boiling phase (made up of the initial
heating and the bubble end point stages) and the boiling phase (made up of the
surface boiling and the falling rate stages)
20
The rate of heat transfer from the oil to the food is controlled by the
surface heat transfer coefficient, h at the boundary between the food and the oil.
Convective heat transfer coefficients for the non-boiling phase have been found
to be in the range of 250-300 W/m2°C (Moreira et. al., 1995a; Miller et. al., 1994).
Values of 300 W/m2°C at the onset of boiling and a peak value of 1100±140
W/m2°C during frying have also been reported (Hubbard and Farkas, 1999).
Baik and Mittal (2002) determined and analyzed h at two locations (top
and bottom surfaces) of a tofu disc during frying based on energy and mass
balance. It was reported, in this study, that the top surface temperatures were
lower than those of the bottom in the beginning of frying (less than 40-600 s) at
147-172°C oil temperature. Later, temperatures at the top surface exceeded
those at the bottom. The authors reported that h for the bottom surface was
higher at the beginning and it was attributed to different magnitudes of natural
convection at the two portions without significant bubble release at the onset of
frying. After vapour release, h at the top surface was higher than that for the
bottom surface for all frying cases. The authors speculated that the higher
magnitude of agitation at the top surface resulted in the higher h value.
Moreira and co-workers (1995a) reported that convective heat transfer
coefficient is dependent on the interaction between the temperature gradient
and the drying rate in a frying process. As the thermal gradient between the oil
and food increases, vapour loss rate also increased which additionally serve to
further agitate the surrounding oil and increase the heat transfer coefficient.
Increase in oil temperature also leads to low oil viscosity and hence its
21
resistance to flow, thereby further enhancing the heat transfer coefficient. It was
reported that h value during frying is affected by bubble flow direction, velocity,
bubble frequency and magnitude of oil agitation (Baik and Mittal, 2002). A
hypothesis linking the increase in heat flux with increasing oil degradation
through a reduction in the vapour bubble size and increase in the bubble
frequency has been suggested (Farkas and Hubbard, 2000).
Tong and Tang (1997) found that formation and departure of vapour
bubbles during boiling enhances heat flux to a material by forced convection of
fluid across the material. However, the presence of bubbles on the food surface
during frying insulates the surface from the heating media and can therefore
reduce heat flux (Farkas and Hubbard, 2000). The effect of bubbling on the h
value increases with the oil temperature during frying of potato. During the
boiling phase, h values are 80% greater than those obtained without boiling for
oil temperature of 180°C, while the increase in h value of boiling phase over
non-boiling phase for oil temperature of 140°C is only 40% (Costa et. al., 1999).
This last observation can be attributed to the fact that higher temperature
reduces the viscosity of the frying oil, which leads to further agitation of the
frying process and thus, an increase in the h value. In this same study, the
researchers also concluded that when moisture loss rate is high, the bubbles
near the food surface may hinder the heat transfer since they serve as insulation
pockets and also that the fraction of heat used for evaporation in a frying
process depends on the temperature gradient between food and oil, water loss
rate and the h value.
22
2.3.3 Moisture Transfer
Moisture loss during frying was expressed by the following general form
(Rice and Gamble, 1989):
Rate of moisture transfer = Driving force/Resistance
(2.14)
The driving force is provided by the conversion of water to steam, which
forces its way out of the product while resistance to mass transfer is provided by
internal resistance to mass diffusion and the surface resistance of the product.
During frying, the free water at the surface of potato chips evaporated rapidly,
the surface becomes dry and the inner moisture is converted to vapour, thereby,
creating a vapour gradient (Rice and Gamble, 1989; Gamble and Rice, 1987,
1988).
Krokida and co-workers (2000) demonstrated that oil temperature has a
negative effect on the moisture content of french-fried potatoes. As frying
temperature increases, the moisture content for the same frying time decreases
since an increase in temperature results in a higher kinetic energy for water
molecules leading to a more rapid moisture loss in the form of vapour which
ultimately reduces the moisture content of the product. Moisture removal rates
were found to be higher in the top portion and at higher temperatures than in the
bottom portion and at lower temperatures, respectively, during frying of a tofu
disc (Baik and Mittal, 2002). Costa and co-workers (1999) used image analysis
of bubbles to estimate water loss rate of potato during frying and observed a
relationship between vapour build-up and water loss rate. Different flow patterns
23
were noticed for bubbles from top, side and bottom surfaces of potato during
frying. Since this will lead to different agitation patterns, it was suggested that
heat transfer coefficient might be expected to be position dependent. Also, there
was a change in potato sample geometry during frying which also changed the
oil agitation patterns and consequently changed the heat transfer coefficient. As
frying proceeded, bubbling was found to increase to a maximum value and then
decreased.
In a study on the effects of initial moisture content on sample temperature,
moisture loss and crust thickness, Ni and Datta (1999) found moisture loss to
increase significantly with initial moisture content because the surface
evaporation and subsequent internal evaporation are much higher for high
moisture food. The researchers also studied the vapour and liquid water fluxes
in the crust and the core regions. For the crust region, it was reported that
vapour diffused from the evaporation front to the product surface, with the
maximum diffusional flux occurring near the evaporation front and its magnitude
decreasing with the frying time since vapour concentration decreases with
moisture content. In the core region, the capillary diffusional flux of water is
towards the surface. There exist regions of constant flux where water saturation
is spatially linear and capillary diffusivity is relatively constant. For the same core
region, the authors also reported existence of water convective flux towards the
center due to pressure. It has a smaller but comparable magnitude with capillary
diffusional flux. It was concluded that neither transport mechanism in the core
region could be ignored. Within oil temperature range of 145°C to 185°C, the
24
diffusion coefficients for moisture loss during frying of potato was found to
increase from 58×10-10 to 109×10-10 m2/s (Rice and Gamble 1989).
2.4 Mathematical modeling
Several mathematical models for frying at different levels of complexity
have been reported. Rice and Gamble (1989) used analytical solution of Fick’s
second law to develop a one-dimensional diffusion model for potato slice frying.
2 2



8 ∞ 
1
c − c1
 exp − (2n + 1) π Dt   
= 2 ∑

x 2   
4
co − c1 π n = 0  (2n + 1)2 

(2.15)
Where:
c = average concentration at any time (kg/m3)
c0 = initial average concentration (kg/m3)
c1 = surface concentration = 0 (kg/m3)
x = slab thickness (m)
D = moisture diffusion coefficient (m2/s)
t = time (s)
Rice and Gamble (1989) assumed that the potato slice is an infinite slab
and ignored the second and subsequent terms for long frying time (Fo > 2.0). A
negligible surface resistance was assumed and relationship between apparent
diffusion coefficient and temperature was also assumed to be an Arrhenius-type.
The authors also related frying time, t (s) to the square of moisture loss (% initial
weight).
25
m2 = at + C
(2.16)
Where the intercept (C) was linearly related to oil temperature T (°C) by the
regression equation:
C = 40T - 6036
(2.17)
Semi empirical relationships were proposed for heat and mass transfer during
frying of crustless and crust-forming products (Mittelman et. al., 1984). The
amount of water evaporated was found to be proportional to the square root of
the frying time and the temperature difference between the oil and boiling water.
For the crustless product, it was assumed that evaporation of water takes place
from a receding front of liquid water at boiling temperature, the mass fraction of
the solid matrix in the wet food is very low and that all the heat transferred from
the oil to the food is used up in evaporation of water. The following equation was
proposed:
m2 =
2kA 2 ρ
(To − Tw )(t − t o )
λ
(2.18)
For the crust-forming product, it was assumed that the temperature of the
core remain constant at 100°C throughout the particle, the thickness of the crust
remains constant throughout the latter stage of frying following the initial
openings and cracks in the crust at the beginning of crust formation. The
following equation was developed:
y 2 = C 2 exp(0.0113Toil )(t − t o )
Where:
y = cumulative quantity of water evaporated (kg)
k = thermal conductivity (W/m°C)
26
(2.19)
A = area of heat and mass transfer (m2)
5 = density of the wet food (practically that of water) (kg/m3)
7 = latent heat of evaporation of water (J/kg)
Toil = temperature of the oil (°C)
Tw = boiling temperature of water (°C)
t = time (s)
to = time elapsed until evaporation starts (s)
C2 = proportionality constant.
Farkas and co-workers (1996a, 1996b) developed a one-dimensional
model for heat and moisture transfer during deep fat frying for material with
infinite slab geometry and a moving boundary for the crust. Separate equations
for the crust and the core regions were given as follow:
For heat transfer, core region:
k effII
∂T
∂T
∂ 2T
+ N βx C pβ
= (ε β ρ β C pβ + ε σ ρ σ C pσ )
2
∂x
∂t
∂x
(2.20)
For heat transfer, crust region:
k effI
∂ 2T
∂T
∂T
= (ε γ ρ γ C pγ + ε σ ρ σ C pσ )
+ N γx C pγ
2
∂t
∂x
∂x
(2.21)
For mass transfer, core region:
∂c β
∂t
= Dβσ
∂ 2cβ
∂x 2
(2.22)
For mass transfer, crust region:
∂  ∂Pγ 
=0
ρ γ
∂x 
∂x 
27
(2.23)
Where:
keff = effective thermal conductivity (W/m.K)
T = temperature (°C)
ε i = volume fraction of specie i ( mi3 / mt3 )
C pi = specific heat of specie i (J/kg.K)
N ix = flux of species i in x-direction ( kg / mt2 s )
P = pressure (N/m2)
t = time (s)
ρ i = density of species i ( kg / mi3 )
c β =24
5
6
4
7
6
8
9
8
5
6
25
2
2
2 ε β ρ β 2
mt3 )
D = mass diffusivity (m2/s)
x = distance in the direction of heat or mass transfer
9
4
4
4
are Liquid water, water vapour and solid phases respectively.
The partial differential equations were solved by a numerical method
using coordinate transformation of the equations, application of the finite
difference method of Crank-Nicholson and finally Gauss-Seidel iteration. The
researchers assumed advection of energy due to oil flux is negligible and that
the mass fraction of oil in the fried material is also negligible. The model did not
include fat transfer. Pressure-driven flow was considered only for the vapor
phase in the crust region. Both diffusional flow in the crust region and pressuredriven flow in the core region were ignored. One-dimensional models with
boundary and initial conditions for moisture and heat transfer for a tortilla chip
28
during frying were derived with diffusion equation by Moreira et. al., (1995a,
1995b), as follows:
Mass balance for moisture content of the product,
∂ 2m 
∂m
= D 2 
∂t
 ∂x 
(2.24)
Energy Balance for the product,
 ∂ 2T 
∂T
=α 2 
∂t
 ∂x 
(2.25)
Moreira et. al. (1995a, 1995b) also described oil uptake as a function of the
frying time by a first order exponential equation,
M
f
(t ) = M fe [1 − exp(− k ′t )]
(2.26)
Where:
m = moisture content (% d.b.)
D = mass diffusivity (m2/s)
x = variable distance along thickness of tortilla chip (m)
Mf = oil content (%)
1 = thermal diffusivity (m2/s)
Mfe = final oil content (%)
k ′ = a constant (1/s)
t = time (s)
T = temperature (°C).
Negligible shrinkage and constant thermal and moisture diffusivities were
assumed. No transport model for the oil phase was provided. The equations
29
were solved using explicit finite difference technique.
Atteba and Mittal (1994) used a one-dimensional model to describe heat,
moisture and fat transfer for deep fat frying of beef meatballs. In the model, a
general diffusion equation was used for heat transfer, moisture transfer and fat
transfer in the absorption period. Capillary flow equation was used for fat
transfer during desorption period. The models developed in non-dimensional
form were:
For moisture transfer,
∂C Dm  2 ∂C ∂ 2 C
+
= 2 
∂t
R  ψ ∂ψ ∂ψ 2

 ;

(2.27)
For fat transfer (absorption),
D f  2 ∂C f ∂ 2 C f
= 2
+
∂t
∂ψ 2
R  ψ ∂ψ
∂C f

 ;


(2.28)
For heat transfer,
∂θ
α  2 ∂θ ∂ 2θ 

 ;
=
+
∂t R 2  ψ ∂ψ ∂ψ 2 
(2.29)
Where:
C=
m − me
;
m o − me
(2.30)
Cf =
mf − mf e
;
mf o − mf e
(2.31)
θ=
T − To
;
Ta − To
(2.32)
30
ψ =
r
.
R
(2.33)
For fat transfer (desorption),
V
dmf
= − KLA[mf − mfep ] ;
dt
(2.34)
Where:
KL = fat conductivity (m/s)
V = product volume (m3)
A = surface area (m2)
mfep = equilibrium fat content (dry basis)
C = concentration
R = radius (m)
r = radial distance (m);
3
4
= dimensionless distance
1 = thermal diffusivity (m2/s)
D = mass diffusivity (m2/s)
t = time (s)
T = temperature (°C)
6 = dimensionless temperature
m = moisture content (dry basis)
Subscripts m, f, e, o, a represent moisture, fat, equilibrium, initial and ambient
respectively.
The equations were solved using finite difference technique. Fat and
water were assumed to be mobilized by concentration gradients. The effect of
31
shrinkage and crust formation on physical properties of the product was also
neglected. No evaporation term was included in the energy or moisture transport
equations of this model except for the inclusion of surface evaporation as
boundary condition for energy equation. The model was able to predict fat
transfer during deep fat frying of a fatty product.
A multiphase porous media model was developed to simulate frying of
potato slices (Ni and Datta, 1999). The model included the significance of
pressure-driven flow for the oil, vapour and air phase in a non-hygroscopic
porous medium. Five conservation equations for water vapour, liquid water, air,
oil and energy in the porous medium were derived:
∂c v
→ .
+ ∇ n v  = I
∂t
 
(2.35)
.
∂c w
→
+ ∇ n w  = − I
∂t
 
(2.36)
∂c a
→
+ ∇ n a  = 0
∂t
 
(2.37)
∂co
→
+ ∇ no  = 0
∂t
 
(2.38)
(ρc )
p eff
.
∂T
= ∇(k eff ∇T ) − λ I
∂t
(2.39)
Where:
(ρc )
p eff
= φ (S g ρ g c pg + S w ρ w c pw + S o ρ o c po ) + (1 − φ )ρ s c ps
k eff = φ (S g k g + S w k w + S o k o ) + (1 − φ )k s
(2.40)
(2.41)
c = mass concentration (kg/m3)
32
n = total flux (kg/m2.s)
t = time (s)
φ = porosity
5 = density (kg/m3)
P = pressure (Pa)
S = saturation
cp = specific heat capacity (J/kg.K)
T = temperature (K)
.
I = volumetric expansion (kg/m3.s)
k = thermal conductivity (W/m.K);
7 = latent heat of vaporization (J/kg).
Subscripts: v, w, a, o, g, s, p, eff are vapour, water, air, oil, gas (vapour+air),
solid matrix (surface), partial pressure, and effective, respectively.
The five governing equations were transformed into four equations with
variables Sw, So, T and P and were solved with a central finite difference method
with initial and boundary conditions. Ni and Datta (1999) assumed the existence
of thermal equilibrium between phases and ignored the contribution of
convection to energy transport. The model was able to consider the transport of
oil, water, vapour and air components separately. However, the model did not
account for changes in porosity and its effect on energy and mass transport.
2.5 Heat and Mass Transfer Analogy
In deep fat frying, heat and mass transfer occur simultaneously. Even
though researchers tend to separate the two processes for the purpose of
33
analysis, a complete analysis of one cannot be done without the other. A host of
scientists has been involved in the historic development of the idea that a similar
phenomenon guides both heat and mass transfer. Attempts have also been
made to estimate one transport property (e.g. D) from another (e.g. k) as found
in the Lewis equation.
Reynolds’ analogy is probably the earliest known of heat and mass
transfer analogies. Reynolds argued that under certain conditions, heat, mass
and momentum transfer occur at the same rate (Incropera and Dewitt, 2004).
The logic for this was that each of these transfer processes involves:
(1) natural diffusion in a fluid at rest; and
(2) eddies which bring fresh fluid into contact with a surface and allows
transfer of specie to the surface.
While the former is independent of velocity, it is obvious that the latter is
not. So, for turbulent flow, which is typical of frying, a mass flux can be
calculated from:
n = hm ∆C = (a + bU )∆C
(2.42)
Similarly, a heat flux can be expressed as:
q = h∆T = (a + b U )∆( ρC p T )
(2.43)
Where:
n = mass flux (kg/m2)
hm = mass transfer coefficient (kg/m2.s)
C = concentration of species (kg)
U = velocity (m/s2)
34
q = heat flux (W/m2)
h = heat transfer coefficient (W/m2.K)
T = temperature (°C)
5 = intrinsic density (kg/m3)
Cp = specific heat capacity (J/kg.K).
a, a', b and b' are constants
For very turbulent flow, it is assumed that a = a' = 0, i.e. turbulent effects
dominate. At this point, Reynolds also assumed that heat and mass transports
were arising from the same turbulent mechanism and therefore assumed that b
= b'. It follows then that the transport properties simplify to:
h m = bU
(2.44)
h
= b′U
ρC p
(2.45)
And since b = b', then
hm
h
=
U ρC pU
(2.46)
Where:
hm = mass transfer coefficient (kg/m2.s)
h = heat transfer coefficient (W/m2°C)
U = velocity (m/s)
5 = density (kg/m3)
Cp = heat capacity (J/kg.K)
The benefit of this work is that we can derive one transfer coefficient if we
know the other. The Reynolds’ analogy is true for gases but not for liquids. This
35
is better understood by recognizing that turbulent flow takes place at two levels:
at an eddy level (macroscopic) and at a diffusional level (microscopic). The latter
is buried in b, b'since Reynolds assume a = a'= 0. For gases, the diffusional
level is expressed by D ≅ α ≅ ν ≅ 0.1 m2/s. Or in a better sense: ν/D ≅ ν/α ≅ 1,
i.e. Sc = Pr. Which makes the analogy sensible.
For liquids, however, Sc and Pr are about 1000 and 10 respectively. (Sc = µ/(ρ
D) and Pr = Cp µ/k). These are definitely not the same and it invalidates the
analogy for liquids.
Where
D = mass diffusivity (m2/s)
v = kinematic viscosity (m2/s)
1 = thermal diffusivity (m2/s)
Pr = Prandtl number (dimensionless)
Sc = Schmidt number (dimensionless)
k = thermal conductivity (W/m.K)
2 = dynamic viscosity (kg/m.s)
Reynolds ignored the laminar boundary layer in his work and a couple of
scientists Taylor and Prandtl working with Reynolds analogy broke the flow into
two regions: a laminar layer and a fully turbulent layer (Brodkey and Hershey,
1988). Using Fourier’s and Newton’s laws across the sub-layer, they developed
a correction factor as below:
h
h
(1 + α (Pr − 1)) = m (1 + α ( Sc − 1))
C p ρU
U
36
(2.47)
It can be seen from the result that for a very turbulent flow such that the
boundary layer approaches zero (14
= 0), then 1/(1+α(Pr-1)) approaches unity
and this result approaches Reynolds analogy.
Chilton-Colburn analogy is an attempt to extend the Reynolds’ analogy to liquids.
It is purely empirical in basis. Chilton-Colburn decided that the basic form of the
Reynolds’ analogy was good, but that the b'
s should be replaced as follows:
2
h
b' =
ρC pU
ν 3
 
α 
(2.48)
2
h  ν 3
b= m 
U D
(2.49)
Such that;
2
2
hm  ν  3
h ν 3
 
  =
U D
ρC pU  α 
(2.50)
Once again, with the knowledge of one transfer coefficient, the other can
be calculated from this equation. Due to the reason stated earlier regarding
variation in Schmidt and Prandtl number for liquids, it is difficult to apply these
results to frying. The best approach for now therefore is still empirical correlation
using experimental results. The disadvantage of this approach is that the models
developed are most suited only to products with similar geometry, physical
properties and frying process requirements as the one used in developing the
model.
37
2.6 Summary
Literature shows sweetpotato is an important crop. A lot of studies have
been conducted on deep fat frying. Most of the works focused on heat transfer
during frying, mechanisms and rate of moisture loss and fat uptake, quality
kinetics of the fried food and the frying medium and mathematical modeling of
the frying process. However information on thermal properties and deep fat
frying of sweetpotato is lacking in literature. Also no work has been reported on
heat and mass transfer analogy during deep fat frying.
38
3. MATERIALS AND METHODS
This chapter presents the materials used in this study and the methods
that were used in conducting experiments or estimating properties values.
3.1 Materials
Sunny II sweetpotatoes (Ipomoea batatas) were bought in 20 kg lots from
Eastern Market, (Saskatoon, SK) to avoid variability in quality and product
chemistry. The potatoes were stored in a cooling chamber at a temperature
range of 11±1°C usually for 24 h before use. The tubers were manually peeled
with a hand peeler and then cut into discs using a cylindrical cutter and a knife.
Samples were cut into discs having diameters of 2.5 cm, 3.5 cm and 4.0 cm, all
with 1 ± 0.1 cm thickness. Canola oil (Sunnyfresh Ltd, Toronto, ON) was used
as the frying oil in this study.
3.2 Density Measurement
Density of a material is the ratio of its mass to the volume it occupies.
Two methods: Gas multipycnometer and a mechanistic model developed by
Choi and Okos (1985) were used in this study.
39
3.2.1 Multipycnometer
Density of sweetpotato was measured at four different levels of moisture
content 0.45, 0.55, 0.65, 0.75 w.b. Moisture content of sweetpotato was
adjusted by heating the sample in a microwave oven (Panasonic NN-5553C,
Matsushita Corporation, Franklin Park, IL). The oven had been previously
calibrated to determine the heating time to get sample to a desired moisture
content level. The microwave oven initially has been calibrated at a constant
power (800 W) to determine the heating time required to dry sample to the
required moisture content. After heating, the sample was equilibrated for 6 h and
the moisture content was determined by oven method of AOAC standard 984.25
(AOAC, 2002) to check if the moisture content is of the expected value.
Density of sweetpotato was determined in 5 replicates using the
multipycnometer (Quantachrome Corporation, Boynton Beach, FL). The
instrument is shown in Figure 3.1. It measures the true volume of a solid
material. It makes use of two cylindrical volumes; reference volume which is
empty and the other volume containing the sample. The instrument makes use
of a displacement fluid with very small atomic dimension to penetrate the pores
of the material.
40
Figure 3.1: Multipycnometer used to measure density of sweetpotato.
The volume of the sample is given by the equation:
 P  
VP = VC − VR  1  − 1
 P2  
(3.1)
Where:
Vp = volume of the sample (cm3);
Vc = volume of the cell containing the sample (cm3);
VR = volume of the reference cell (cm3);
P1 = pressure reading after pressurizing the reference volume (psi); and
P2 = pressure reading after including Vc (psi).
True density of the sample was then determined from the ratio of sample
mass to its volume determined above.
41
ρ=
m
VP
(3.2)
Where:
5 = density (kg/m3)
VP = volume of sample (m3) as determined from multipycnometer
m = mass (kg).
Sample mass was determined by weighing the sample on an Ohaus
GA2000 balance (Ohaus Corp., Pine Brook, NJ). Multipycnometer was
calibrated before use as suggested in the manual to ensure that a correct value
for the reference volume was used. A regression model was developed with
SAS (SAS Institute Cary, NC) for the density result from this experiment as a
function of moisture content.
3.2.2 Mechanistic Model
Choi and Okos (1985) developed models for thermal properties of food
products. Models were developed for estimating properties like density, specific
heat capacity, thermal conductivity and thermal diffusivity of major components
of food: carbohydrate, fat, protein, ash, fibre and water. Percentage composition
of these major components in sweetpotato was derived from USDA agricultural
handbook (USDA, 1963). Choi and Okos (1985) have demonstrated that thermal
properties of food materials are dependent on composition of major food
components in the particular food. Variation of these major food components
among different samples of a food variety is negligible. The models give very
accurate predictions for a wide range of food materials. Also, solid components
42
were shown to have lesser influence than that of water on the thermal properties
of the food. It was on this basis that the developed mechanistic models for major
food components (Choi and Okos, 1985) and food composition data for
sweetpotato (USDA, 1963) were used in this study.
The models were developed as a function of temperature and are
presented in Table 3.1. Percentage composition of major food components in
sweetpotato is presented in Table 3.2. The particular property for a food product
is then estimated by combining the results of models in Table 3.1 with mass
fraction (specific heat and density) or volume fraction (thermal conductivity and
thermal diffusivity) of each component using series or parallel models. The
model for density (equation 3.3) was used in this study to estimate the density of
sweetpotato both as a function of temperature and moisture content. Based on
the chemical composition of sweetpotato, Excel (Microsoft Corp. Redmond WA)
was used to generate density with moisture content range of 0.45-0.75 w.b. and
temperature range of 20-150°C. The series model was chosen over the parallel
model for estimating density because density variation among major food
components is wide. Density of water and fat are different from densities of other
components. The series model predicts a better value in such circumstances
(Stroshine, 1998). A regression equation was developed for the data.
43
Table 3.1. Models of thermal properties and density of major components of
foods. (adapted from Choi and Okos, 1985).
Thermal
Major
Group models
Property
component
temperature function
k
W/m°C
122
m2/s
kJ/kg.K
% error
0.012
5.91
Fat
k = 1.8071 x 10-1 - 2.7604 x 10-3T - 1.7749 x 10-'T2
0.0032
1.95
Carbohydrate k = 2.0141 x 10-1 + 1.3874 x 10-3T - 4.3312 x 10-6T2
0.0134
5.42
Fiber
k = 1.8331x 10-1 + 1.2497 x 10-3T - 3.168 3 x 10-6T2
0.0127
5.55
Ash
k = 3.2962 x 10-1 + 1.4011 x 10-3T - 2.906 9 x 10-6T2
0.0083
2.15
Protein
122= 6.8714 x 10-2 + 4.7578 x 10-4T - 1.4646 x 10-6 T2
0.0038
4.50
Fat
1223
29.8777 x 10-2 - 1.2569 x 10-4T - 3.8286 x 10-8T2
0.0020
2.15
Carbohydrate
1223
28.0842 x 10-2 + 5.3052 x 10-4T - 2.3218 x 10-6T2
0.0058
5.84
Fiber 222222222221223
27.3976 x 10-2 + 5.1902 x 10-4T - 2.2202 x 10-6T2
0.0026
3.14
1223
21.2461 x 10-1 + 3.7321 x 10-4T - 1.2244 x 10-6T2
0.0022
1.6
39.9501
3.07
42= 9.2559 x 102 - 4.1757 x 10-1T
4.2554
0.47
Carbohydrate 2222242= 1.5991 x 103 - 3.1046 x 10-1T
93.1249
5.98
Fiber
242= 1.3115 x 103 - 3.6589 x 10-1T
8.2687
0.64
Ash
42= 2.4238 x 103 - 2.8063 x l0-IT
2.2315
0.09
Protein
Cp = 2.0082 + 1.2089 x 10-3T - 1.3129 x 10 -6T2
0.1147
5.57
Fat
Cp = 1.9842 + 1.4733 x 10-3T - 4.8008 x 10-6T2
0.0236
1.16
Carbohydrate
Cp = 1.5488 + 1.9625 x 10-3T - 5.9399 x 10-6T2
0.0986
5.96
Fiber
Cp = 1.8459 + 1.8306 x 10-3T - 4.650 9 x 10-6T2
0.0293
1.66
Ash
Cp = 1.0926 + 1.8896 x 10-3T - 3.6817 x 10-6T2
0.0296
2.47
Fat
Cp
error
k = 1.7881 x 10-1 + 1.1958 x 10-3T- 2.7178 x 10-6T2
Protein 222222222242= 1.3299 x 103 - 5.1840 x 10-1T
kg/m3
Standard
Protein
Ash
4
Standard
44
Table 3.2. Percentage composition of major food components in sweetpotato
(adapted from USDA, 1963)
Component
Percentage by mass (%)
Protein
1.7
Fat
0.4
Carbohydrate
ρ=
26.3
Fibre
0.7
Ash
1.0
Water
69.9
1
X iw
∑ρ
i
(3.3)
Where:
Xiw = weight fraction of component I (kg)
5i = density of pure component I (kg/m3).
54
= density of sweetpotato (kg/m3)
3.3 Specific heat
Specific heat capacity, Cp is defined as the amount of heat needed to
raise the temperature of 1 kg of a material by 1 degree K. It depends mainly on
the composition of the material, temperature and pressure. Cp also decreases
with a decrease in moisture content. Specific heat capacity for this study was
determined both from mechanistic models and experimentally using a differential
scanning calorimetry (DSC). Specific heat capacity was determined as a
45
function of product temperature (20-180°C) and moisture content (0.45, 0.55,
0.65, 0.75 (w.b)).
3.3.1 Mechanistic model
Specific heat of sweetpotato was estimated as a function of temperature and
moisture content based on the mass fraction of the components using Excel.
The parallel model (equation 3.4) was used since it applies better to non-fibrous
materials where thermal properties are not dependent on direction of heat flow
and variation in Cp of major food components is not large (Stroshine, 1998). A
regression equation was then developed for the generated data using SAS (SAS
Institute, Cary NC).
C p = ∑ C pi X iw
(3.4)
Where:
Xiw = weight fraction of component i
Cpi = specific heat of pure component i (J/kg.K)
Cp = specific heat of sweetpotato (J/kg.K)
3.3.2 Differential Scanning Calorimeter
DSC method is very popular in determining specific heat of food
materials. The technique is direct, relatively quick and dynamic over a wide
range of temperature. However, it is expensive, requires calibration since it is a
comparative device, and requires only a very small product sample which might
make it difficult to obtain homogenous samples that are truly representative of
46
the product.
3.3.2.1 Sample preparation
Sweetpotato samples to be run on DSC were prepared by peeling and
cutting the tuber into tiny pieces of approximately 2 × 2 × 2 mm by surgical blade.
This was by no means precision cutting. The aim was just to have tiny pieces
small enough such that a few pieces would give a mass of about 10-12g
required for the experiment and fit into the sample pan. The sample was then
heated in a microwave oven to get it to desired moisture content. Samples for
this experiment were at four moisture content values of 0.45, 0.55, 0.65 and
0.75 wet bases. The microwave oven initially has been calibrated at a constant
power to determine the heating time required to dry sample to the required
moisture content. After heating, the sample was equilibrated for 6 h and the
moisture content was determined by oven method of AOAC standard 984.25
(AOAC, 2002) to check if the moisture content is of the expected value.
3.3.2.2 Experimental procedure
Differential scanning calorimeter (DSC) shown in Figure 3.2 was also
used in determining the specific heat of sweetpotato. The model is DSC 2010
(TA Instrument Inc., New Castle, DE). The DSC includes a holder housing two
discs which are in thermal contact with each other and are isolated from the
environment. Two sample pans were prepared; one of the pans contained the
sample to be measured while the other was an empty reference pan. The
sample pan contained 10-12 mg of the sample. This was hermetically sealed.
47
The pans were placed on the discs in the holder and the holder was closed. The
discs sat on a raised platform on a constantan disc. Both pans were heated at a
controlled, known heating rate of 10°C/min. in this case and the heat flow
between the pans, which gives the difference in heat capacity of the reference
pan and the sample, was monitored by thermocouples beneath the disc and
measured. Specific heat of the sample is estimated from the heat flow rate,
heating rate and the mass of the sample (equation 3.5). The experiment was
carried out in duplicates. The DSC is a comparative device and must be
calibrated. Prior to use, the DSC was calibrated with water. The data analysis
software used was TA Universal Analysis (TA Instruments, New Castle, DE).
The operational equation of the software is as shown in equation (3.5)
Figure 3.2: DSC 2010 used for specific heat measurement of sweetpotato.
Cp =
dQ / dt
mdT / dt
(3.5)
48
Where:
Cp = heat capacity (J/g.K)
dQ/dt = heat flow rate (J/s)
dT/dt = heating rate (K/s)
m = mass of the sample (g)
Often during the experiment there was leakage (most especially at very
high temperatures) in which moisture vapour escaped from the sample pan. This
is due to increased vapour pressure as heating progresses, if the pans were not
sealed properly. It is therefore important that the pan is properly sealed to avoid
moisture vapour loss during heating. Some of the runs in this study leaked and
they were repeated.
3.4 Thermal Conductivity
Thermal conductivity, k gives the rate at which heat is conducted through
a unit thickness of a material when a unit temperature gradient exists across the
thickness. From Fourier equation, heat flow in the material is given by:
q = −k
∂T
∂x
(3.6)
Where
q = heat flow (W)
k = thermal conductivity (W/m.K)
∂T = change in temperature (K)
∂x = material thickness (m)
49
It is a function of heat flow, area perpendicular to its direction and
temperature drop in a sample. For a porous media consisting of solid and gas
phases, the measured thermal conductivity is an apparent one, usually called
the effective thermal conductivity (keff). It is an overall transport property
assuming that heat is transferred by conduction through the solid and the porous
phase.
Thermal conductivity in this study was determined with mechanistic
models of food major components using the work of Choi and Okos (1985) as
outlined in Table 3.1. Thermal conductivity of sweetpotato was then estimated
from these models as a function of temperature and moisture content based on
the volume fraction of the components (equation 2.6) using Excel. The parallel
model was chosen over the series model to determine thermal conductivity
since thermal conductivity in vegetables are not dependent on the direction of
heat flow as it is in fibrous materials like fish and meat. Also, variation among k
of major food components is not large (Stroshine, 1998). Result from parallel
model also compare better with values in literature. A regression equation was
also developed for the generated data using SAS.
An attempt was made at determining thermal conductivity experimentally
using both the line heat source method and KD2 thermal properties analyzer
(Decagon Devices Inc. Pullman, WA). However, there was significant error in
the results in particular at temperature beyond boiling point (100°C) for the line
heat source method. Some line heat source methods can handle higher
temperature than 100°C, however it would cause significant deviation due to
condensation and moisture loss at higher temperature. Since frying is done at
50
temperature range of 150-170°C, this questions the ability of the line heat
source method to determine thermal conductivity at typical frying temperature.
For the KD2 thermal properties analyzer, its operating temperature is limited to
60°C. This limits its use for studying thermal properties for frying application.
3.5 Thermal diffusivity
Thermal diffusivity is the rate at which heat diffuses through a material.
The property is needed in establishing a temperature history of a body under
transient condition from Fourier’s law of heat conduction. Thermal diffusivity of
sweetpotato in this study was determined with thermal diffusivity models of food
major components using the work of Choi and Okos (1985) shown in Table 3.1.
Thermal diffusivity estimation was based on volume fraction of the components
(equation 3.7). A regression equation was also developed for the generated data
using SAS.
α = ∑ α i X iv
(3.7)
Where:
Xiv = volume fraction of component i (m3)
1i = thermal diffusivity of component i (m2/s)
1 = thermal diffusivity of sweetpotato (m2/s).
3.6 Moisture diffusivity
Moisture diffusivity, D is the rate at which moisture diffuses through a
material. Most of the available methods of determining diffusivity are based on
51
Fick’s laws of diffusion. D in this study was estimated from the relationship in
equation (3.8)
D=
km
ρC m
(3.8)
Where:
D = moisture diffusivity (m2/s)
km = moisture conductivity (kgmoisture/m.s)
5 = density of sample (kg/m3)
Cm = specific moisture capacity (kgmoisture/kgsample)
km and Cm were derived from literature (Sheerlinck et.al., 1996; Chen et.al.,
1999). Density was experimentally determined as described under section 3.2. Since km
and Cm are constants and density was modeled as a function of moisture content only,
moisture diffusivity variation with frying condition is limited to change in moisture
content only. A model that includes the effect of temperature would however have been
more accurate in predicting mass transfer rate.
3.7 Deep fat frying
This is the actual cooking of the sweetpotato in oil. It is usually done at a
temperature above the boiling point of water typically 150-190°C.
3.7.1 Sample preparation
Sweetpotato tubers were peeled with a hand peeler and then cut into
discs using a cylindrical borer and a knife. Samples were cut into discs having
diameters 2.5 cm, 3.5 cm and 4.0 cm, all with 1 ± 0.1 cm thickness (Figure 3.3).
52
Figure 3.3: Sweetpotato sample sizes and cylindrical borers used to cut samples.
3.7.2 Sample holder
Frying is a very turbulent process in which the product moves around
randomly in the oil. One of the major challenges of studying the process is in
making the product stable enough so as to be able to measure its temperature.
For this study, a sample holder (Figure 3.4) was designed and fabricated. It
essentially has a handle made of Teflon which holds the product and the steel
frame that ensures product’s stability in the turbulent oil. One of the handles was
adjustable such that the holder can handle multiple sizes. Three holes were
53
bored into the other handle to accommodate thermal probes which measure the
temperature within the sample (Figure 3.4). Making the arm of a poor conductor
of heat like Teflon ensures heat conduction to sample is limited to oil only and
does not involve heating from the sample holder arm.
Figure 3.4: Sample holder used for frying with thermocouples attached.
3.7.3 Experimental procedure
A domestic deep-fryer (Cool Touch deep fryer, General Electric,
Mississauga, ON) shown in Figure 3.5 was used for this study. The fryer has a
power rating of 1500W, frying volume capacity of 2 litres and regulates frying
temperature to ±1°C. The fryer has a built-in thermostat for oil temperature
regulation. It also has a temperature indicator but a type-K thermocouple was
also used to measure oil temperature during frying. The sample disc was fixed
inside the sample holder and the sample holder was adjusted with a tight-screw
54
to ensure sample was tightly held. Three thermal probes made from type-T
thermocouple were inserted into the sample through the holes in the fixed arm of
the holder and two other type-T thermocouple were tightly pressed to the
surface of the product by a couple of alligator clips built into the sample holder
(Figure 3.4).
Two litres of canola oil was poured into the deep fryer up to the maximum
mark. Fryer was switched on and its temperature was set to the desired level
(150, 160, 170 or 180°C). After the oil temperature has reached the desired
level as indicated by the fryer indicator and the thermocouple, the sample holder
with the sample in it was lifted with a long holder and placed inside the fryer
(Figure 3.6). The sample was kept in the frying oil for the required frying time
which was typically 300 s. A timer was used to regulate the frying period. The
sample was then removed from the oil with the same long holder after the
desired frying time. The thermocouples, most especially the ones on the surface
were checked after each frying period to make sure that they were still in place
and measuring the right temperature. Data from experiments in which the
surface thermocouples were found to be embedded in the sample crust or found
separated from the surface were discarded. Four frying temperatures of 150,
160, 170 and 180°C were used in this study. Also, three sample sizes having
diameters 2.5 cm, 3.5 cm and 4.0 cm, all with 1 ± 0.1 cm thickness were used.
55
Figure 3.5: Deep fryer used for sweetpotato frying.
Figure 3.6: Sweetpotato sample in the oil during frying.
56
3.7.4 Temperature measurement
A CR10X datalogger (Campbell Scientific Instrument, Logan UT) was
used to acquire real time temperature data in this study. Three type-T thermal
probes were used to measure temperature inside the sample while 2 type-T
thermocouples were used to measure the surface temperature of the sample.
The 5 thermocouples were connected to the datalogger and the datalogger was
programmed to obtain temperature data from the sample.
3.7.5 Moisture content measurement
Moisture content was measured in the crust and core parts of the fried
sample and the raw whole sweetpotato sample. Moisture content during frying
was measured at seven time intervals (0, 30, 60, 120, 180, 240 and 300 s) using
the oven method according to AOAC Method 984.25 (AOAC, 2002). Frying was
done for each specific time interval enumerated above; the product was then
removed from the oil and oil on the surface was dried using a blotting paper. For
the measurement of moisture content of whole sample, the sample was cut into
smaller pieces using a surgical knife as mentioned under section 3.3.2.1,
weighed on a mass balance (Ohaus GA2000, Ohaus Corp. PineBrook, NJ) and
then dried in the oven (Blue M, General Signal, Blue Island, IL) at 103ºC for 48 h
and moisture content was determined using the AOAC official method stated
under section 3.3.2.1.
For crust and core part moisture content determination, a surgical blade
was used to carefully remove the core from the crust (Figure 3.7). The part of
interest was then weighed and oven dried to determine the moisture content.
57
Crust
Core
Figure 3.7: Crust and core samples of sweetpotato after frying.
3.8 Heat transfer coefficient
The rate of heat transfer from the oil to the sweetpotato is controlled by
the convective heat transfer coefficient, h at the boundary between the food and
the oil. The heat transfer coefficient, in this study, was estimated from the heat
energy balance during frying. The energy balance during frying equates total
heat transferred by convection from oil to sweetpotato to the sum of energy
spent on heating sweet potato and energy spent on water evaporation. This is
represented by equation (3.9).
hA (T ∞ − T s ) = MC p
dT
dm
+λ
dt
dt
(3.9)
Where:
58
h = heat transfer coefficient (W/m2°C)
A = area (m2)
T = oil temperature (°C)
Ts = sweetpotato’s surface temperature (°C)
M = mass of sweetpotato sample (kg)
Cp = specific heat capacity (J/kg°C)
T = volume temperature of sweetpotato (°C)
t = time (s)
m = moisture content of sweetpotato (kg)
7 = latent heat of evaporation (J/kg)
The frying time (300 s) was divided into 6 periods and average heat
transfer coefficient, h of sweetpotato was calculated for each of these periods
using equation (3.9). A mathematical expression previously developed for
specific heat capacity of sweetpotato as a function of temperature and moisture
content was substituted in equation (3.9) towards determining h. Volume
average temperature was determined by numerical integration of temperature
data at 5 different locations in the sample.
3.9 Computer simulation of heat and mass transfer
This section discusses the method adopted for computer simulation of
heat and moisture transfer in sweetpotato during frying.
3.9.1 COMSOL™ software
Mass transfer coefficient, hm of sweet potato during frying was
59
determined using a computer simulation software (COMSOL™ Multiphysics,
COMSOL Inc. Los Angeles, CA). COMSOL™ is a PDE-based multiphysics tool
that makes use of finite element modeling (FEM). In FEM, a domain defining a
continuum is discretized into simple geometric shapes called elements (Figure
3.8). Properties and the governing relationships are assumed over these
elements and expressed mathematically in terms of unknown values at specific
points in the elements called nodes. The elements in the domain are linked by
an assembly process. Solution of the governing equations of the phenomenon,
initial conditions and boundary conditions in the domain gives the approximate
predictions of the process in the domain (Baik, 2004).
Figure 3.8: Mesh generation in the sweetpotato sample domain.
COMSOL™ has a host of built-in specialized modules for a variety of
field-specific problems. COMSOL™ also has the capability of creating personal
equation-based models by the user. All these can be achieved through the use
of the software’s graphical user interface (GUI) or through the MATLAB™ (The
Mathworks Inc. Natick, MA) prompt.
60
3.9.2 Backgrounds, assumptions and approach to simulation
Two different phenomena, heat transfer and mass transfer were coupled
in this mathematical modeling. Both of the phenomena have an effect on each
other and occur simultaneously during frying. Due to the symmetrical nature of
the sweetpotato sample discs (Figure 3.9), simulation was done in 2-dimension
with the following assumptions:
1. initial temperature and moisture content distribution in sweetpotato is
uniform;
2. the temperature and moisture content fields on the inner boundaries are
symmetrical;
3. the product is homogeneous and isotropic;
4. product shrinkage during frying is negligible;
5. crust was assumed to be negligible and have same properties as whole
sample;
6. moisture movement is by diffusion and moisture diffusivity encompasses
other mechanisms including convection; and
7. a microscopically uniform porous medium was formed after frying and
most oil diffuses into the product during the cooling period.
61
Figure 3.9: Sweetpotato sample disc geometry.
The input data for this simulation were the thermal and physical
properties models of sweetpotato (density, thermal conductivity, specific heat,
moisture diffusivity, heat transfer coefficient and latent heat of evaporation of
moisture) along with experimental parameters like oil temperature, initial product
temperature of sample and physical model of sample. Latent heat of evaporation
for moisture was assumed constant for temperature above 100°C and taken to
be 2.257 × 106 J/kg (Incropera and Dewitt, 1994).
Heat transfer coefficient of sweetpotato determined under section 3.8 was
an input into the model in interpolation form as a function of time. A guess value
for mass transfer coefficient, hm was then an input in a time-defined interpolation
form for the first time step (0-30 s of frying) and process is simulated.
Temperature profile of the sample (product surface and center) and moisture
content profile (surface and whole sample) were extracted and compared to
experimental data under the same set of conditions. hm was adjusted
appropriately until the best fit was obtained between temperature and moisture
content data from simulation and experiment. The simulation process is then
62
repeated for the next time step of 30-60 s of frying while keeping the hm from the
last step (0 - 30 s) as an input for the previous step to guide the present step.
This process was repeated until hm for the last time step (240–300 s) was
obtained, each time comparing data (temperature and moisture content) from
the simulation to experimental data to obtain hm that gives the best fit.
3.9.3 Criterion for best fit
In the simulation, two temperatures (sample surface temperature, Ts and
center temperature, Tc were compared for both experiment and simulation. Also
compared were moisture content of the surface, Ms and that of the centre, Mc for
experiment and simulation. The goal was to minimize the deviation between
experimental data and simulation data and this was done using the root square
deviation of normalized temperature and moisture content data. Best hm data
was adopted from the one that gives the least value of Dev. As given in equation
3.10.
2
2
 T
  ci ,measured − Tci , simulated  +  Tsi ,measured − Tsi , simulated 
 

n 
Tci ,measured
Tsi ,measured

 


Dev = ∑
2
i =1   M
− M c , simulated   M s , measured − M s , simulated
 +  ci ,measured
 +
 
 
M
M si , simulated
ci
measured
,
 
 




2

 
 
 
n
(3.10)
3.9.4 Governing equations, initial conditions and boundary conditions
Computer simulation of frying was done with COMSOL™ multiphysics
using governing equations for heat and mass transfer with initial and boundary
conditions on a domain representing the product.
63
3.9.4.1 Governing equations
The governing differential equation describing temperature change in the
sweetpotato disc during frying is described in equation (3.11) (Incropera and
Dewitt, 1994; Moriera et. al., 1999). The first term on the left represents axial
spatial temperature change in the domain, the second term on the left represent
radial spatial temperature change in the domain, the third term on the left
represent 8
7
2
7
8
28
9
6
7
9
24
7
2
2
5
6
25
2
8
7
9
2
5
9
2
7
9
7
2
2
28
7
2
water vapour flux. The right hand term represents temperature change with time.
∂  ∂T  1 ∂  ∂T  ∂ (ΓC pwT ) ∂ (ρC pT )
=
 kr
k
−
+
∂x  ∂x  r ∂r  ∂r 
∂x
∂t
Γ=−
∂ (ρDM )
∂x
(3.11)
(3.12)
Equation (3.13) describes moisture transfer during frying. The left hand term
represents spatial moisture content change in the domain while the right hand
term represents moisture change with time.
∂
∂M
 Dρ
∂x 
∂x
∂M
 1 ∂
 Dρr
+
∂r
 r ∂r 
 ∂ ( ρM )
=
∂t

(3.13)
3.9.4.2 Initial and boundary conditions
Initial conditions: Temperature and moisture content within the
sweetpotato sample prior to frying are uniform and equal to the determined initial
values.
T (x,0) = T0 ;
T (y,0) = T0
(3.14)
M (x,0) = M0 ;
M (y,0) = M0
(3.15)
64
Boundary conditions:
At the centre: For the symmetrical domain in the model, rate of temperature
change and rate of moisture content change at the sample centre is equal to
zero,
∂T
=0
∂x
;
∂M
∂T
∂M
=0 ;
=0 ;
=0
∂x
∂r
∂r
(3.16)
At the surface: At any time, energy transferred by convection from the oil
to the product surface is equal to energy required for transferring heat to the
centre of the product by conduction, for evaporating water from the product and
for heating the water vapour evaporated from the product to oil temperature.
h(T sur − Toil ) = −k
∂T k ∂ (rT )
+ λΓ + ΓC pw (T sur − Toil )
−
∂x r ∂r
(3.17)
Rate of moisture diffusion by vapour flux within the product is equal to
convective moisture transfer rate from the product surface to the oil.
hm ρ (M sur − M ω ) =
∂ ( ρDM ) 1 ∂ (rρDM )
+
∂x
r
∂r
Where:
T = temperature (°C)
t = time (s)
x = axial distance in sample
y = radial distance in sample
r = radial distance in sample
k = thermal conductivity (W/m.K)
Cp = heat capacity of sample (J/kg.K)
Cpw = heat capacity of water (J/kg.K)
65
(3.19)
5 = density (kg/m3)
D = moisture diffusivity (m2/s)
M = moisture content in (kg)
7 = heat of vaporization (J/kg)
h = heat transfer coefficient (W/m2.K)
hm = mass transfer coefficient (kg/m2.s)
Subscripts sur and oil means surface and frying oil respectively.
The equations, initial conditions and boundary conditions were input into
the simulation software and sweetpotato frying was simulated at typical frying
conditions as discussed above.
66
4. RESULTS AND DISCUSSIONS
4.1 Density
Density was obtained in two different ways, multipycnometer and a
mechanistic model.
4.1.1 Measured density
Density of sweetpotato was determined and modeled as a function of
moisture content. Four moisture content levels of 0.75, 0.65, 0.55 and 0.45 were
used. The result from the experiment is presented in Figure 4.1 and Table A1.
Density was modeled as a function of moisture content (w.b) (equation 4.1).
Coefficient of determination (R2) was 0.97 and mean square error (MSE) was
156.1. For typical frying moisture content range, density was 1093-1203 kg/m3.
67
1600
3
Average density (kg/m )
1400
1200
1000
800
600
400
200
0
0.3
0.4
0.5
0.6
Moisture content (w.b)
0.7
0.8
Figure 4.1: Average density of sweetpotato as a function of moisture content at
95% confidence level determined using a multipycnometer.
5 = 1553 – 568.8m
(4.1)
4.1.2 Density from mechanistic model
Density estimates from the mechanistic models yielded a result close to
that of multipycnometer when average value for a moisture content level was
compared. Mechanistic models developed for density of individual major
components of food as a function of temperature, given in Table 3.1, was used
to develop an empirical correlation predicting density of sweetpotato as a
function of temperature and moisture content. The models in Table 3.1 give the
contribution of each food component to density of sweetpotato at a particular
temperature. Moisture content variation was simulated by varying the
percentage of water in equation (3.3), which already includes the effect the
68
effect of temperature variation from the mechanistic models, using Excel.
Density data was generated for temperature for temperature range of 20 to
180°C in steps of 10°C and moisture content range of 40 to 70% w.b. in steps of
1%. The regression model developed from the data generated using stepwise
selection of variables (SAS Institute, Cary NC) is shown in equation (4.2). Direct
comparison is impossible since density in multipycnometer was modeled as a
function of moisture content only. A 3D mesh plot from SigmaPlot (Systat
Software Inc, Richmond CA) of density as a function of temperature and
moisture content is shown in Figure 4.2. Density of sweetpotato reduces with
increase in both moisture content and temperature. This is attributed to the fact
that an increase in temperature leads to higher moisture loss rate and shrinkage
in sample leading to a lower volume. Although mass changes too, the rate of
mass change (reduction) was lower than volume change. Predicted density
range from the model for typical frying conditions (temperature and moisture
content) was 1082-1193 kg/m3. For sweetpotato at high moisture content,
density is close to that of water.
5 = 1491.5 - 0.573T - 496.8m
Where:
54
4
density (kg/m3)
T = temperature (°C)
m = moisture content (w.b.)
69
(4.2)
1300
3)
Density (kg/m
1250
950
1000
1050
1100
1150
1200
1250
1300
1200
1150
1100
1050
1000
Mo 950
ist 0.50
ure
co 0.55
n te
nt 0.60
(w
et
ba
s
20
40
60
80
100
120
0.65
is)
u re
rat
e
mp
)
(°C
140
0.70
160
180
Te
Figure 4.2: Density of sweetpotato as a function of moisture content and
temperature determined from mechanistic model.
4.2 Specific heat capacity
Specific heat capacity, Cp of sweetpotato in this study was obtained by
two methods namely differential scanning calorimeter (DSC) and mechanistic
models of Choi and Okos (1985) as discussed in chapter 3. Results obtained
from both methods are discussed below.
4.2.1 Specific heat from differential scanning calorimeter method
Specific heat was determined as a function of moisture content and
temperature. Four levels of moisture content; 0.45, 0.55, 0.65 and 0.75 wet
basis was used and scanning was done at 10°C/minute between 20 and 180°C.
Experiment was conducted in duplicates. DSC is a comparison device. This
70
means that there is need to calibrate the device with a material of known specific
heat value. The DSC was therefore first calibrated with water prior to the
experiment on sweetpotato.
The result from the experiment is shown in Table A2. Figure 4.3 shows a
3D mesh plot of average Cp values from the experiment as a function of
temperature and moisture content. It can be seen from the plot that Cp of sweet
potato increases with both temperature and moisture content. Specific heat
value obtained from the experiment ranged from 2250-3550 J/kg°C. The
empirical correlation developed using stepwise selection of variables (SAS
Institute, Cary NC) is shown in equation (4.3). R2 was 0.97 and MSE was 3204.5.
Predicted Cp range from the model for typical frying conditions (temperature
range of 150 -180°C and moisture content range 0.45 -0.75(w.b)) was 31003250 J/kg°C.
Cp = 489.8 + 3313m + 24T – 53mT + 33.7m2T
Where:
Cp = specific heat (J/kg°C)
T = temperature (°C)
m = moisture content (w.b.)
71
(4.3)
2200
2400
2600
2800
3000
3200
3400
3600
3800
pacity (J/kg°C)
Specific heat ca
3800
3600
3400
3200
3000
2800
180
160
140
120
100
80
60
(°C
)
2600
a tu
re
2400
tu r e
mp
0.65
Mois
0.60
c o n te
40
0.55
nt ( w
0.50
et ba
0.45
Te
0.70
er
2200
20
sis)
Figure 4.3: Specific heat capacity of sweetpotato as a function of temperature
and moisture content (from DSC).
4.2.2 Specific heat from mechanistic model
Specific heat derived using developed models based on the work of Choi
and Okos (1985) yielded a result close to that of DSC. Mechanistic models
developed for specific heat of individual major components of food as a function
of temperature, given in Table 3.1, was used to develop a model predicting
specific heat of sweetpotato as a function of temperature and moisture content.
The models in Table 3.1 give the contribution of each food component to
specific heat of sweetpotato at a particular temperature. Moisture content
variation was simulated by varying the percentage of water in equation (3.4),
72
which already included the effect of temperature variation from the mechanistic
models, using Excel. Specific heat data was generated for temperature range of
20 to 180°C in steps of 10°C and moisture content range of 40 to 70% w.b. in
steps of 1%. Specific heat increased linearly with increase in both temperature
and moisture content.
A 3D mesh plot of average Cp value from the mechanistic model as a
function of temperature and moisture content is shown in Figure 4.4. The
regression model developed using stepwise selection of variables is shown in
equation (4.4). Predicted Cp range from the model for typical frying conditions
(temperature and moisture content) was 3145-3320 J/kg°C. Moreira and coworkers (1995) reported Cp of 2560 - 3360 J/kg.K for tortilla chip during frying.
Rice and co-workers (1988) reported Cp of 2531 – 4015 J/kg.K for potato.
Cp = 1602.4 + 2543.2m + 0.92T – 9.8×10-5T2
Where:
Cp = specific heat (J/kg°C)
T = temperature (°C)
m = moisture content (w.b.)
73
(4.4)
pacity (J/kg/°C)
Specific heat ca
3600
2600
2800
3000
3200
3400
3600
3400
3200
180
160
140
120
100
80
60
2600
0.65
M oi s
0.60
0.55
ture C
o n te n
40
0.50
t (we
t bas
0.45
Te
m
pe
ra
tu
re
(°C
2800
)
3000
20
is)
Figure 4.4: Specific heat capacity of sweetpotato as a function of temperature
and moisture content (from mechanistic model).
4.3 Thermal conductivity
Thermal conductivity, k of sweetpotato determined from mechanistic
model of Choi and Okos (1985) was a function of moisture content and
temperature. Mechanistic models developed for thermal conductivity of
individual major components of food as a function of temperature, given in Table
3.1, was used to develop a model predicting thermal conductivity of sweetpotato
as a function of temperature and moisture content. The thermal conductivity
models in Table 3.1 give the contribution of each food component by volume
fraction to thermal conductivity of sweetpotato at a particular temperature.
Moisture content variation was simulated by varying the percentage by volume
74
of water in equation (2.6), which already includes the effect the effect of
temperature variation from the mechanistic models, using Excel. Thermal
conductivity data was generated for temperature range of 20 to 180°C in steps
of 10°C and moisture content range of 40 to 70% w.b. in steps of 1%. Thermal
conductivity estimated with this method increased with increase in temperature
and moisture content. Figure 4.5 shows a 3D mesh plot of thermal conductivity
of sweet potato. The regression model developed is shown in equation (4.5).
Predicted k range from the model for typical frying conditions (temperature and
moisture content) is 0.43-0.54 W/m°C. The values are close to reported k values
for similar product in literature. Buhri and Singh (1993) reported thermal
conductivity of 0.552, 0.564 and 0.405 W/m°C for potato, carrot and green apple
respectively at temperature range of 40-50°C.
k = 0.1613 + 0.0014T + 0.2924m + 3.4839 × 10-7T2
Where:
k = Thermal conductivity (W/m°C)
T = temperature (°C)
m = moisture content (w.b.)
75
(4.5)
(W/m.°C)
onductivity
C
l
a
rm
e
h
T
0.65
0.25
0.30
0.35
0.40
0.45
0.50
0.55
0.60
0.65
0.60
0.55
0.50
0.45
0.40
0.35
0.30
0.70
0.25
0.65
160
140
0.60
120
Te 100
mp 80
era
60
tur
e ( 40
°C
)
0.55
0.50
20
0.45
oi
M
re
stu
Co
en
nt
t(
we
as
tb
is)
Figure 4.5: Thermal conductivity of sweetpotato as a function of moisture
content and temperature (from mechanistic model).
4.4 Thermal diffusivity
Thermal diffusivity, 1 of sweet potato was determined as a function of
moisture content and temperature from mechanistic model of major food
components. Mechanistic models developed for thermal diffusivity of individual
major components of food as a function of temperature, given in Table 3.1, was
used to develop a model predicting thermal diffusivity of sweetpotato as a
function of temperature and moisture content. The thermal diffusivity models in
Table 3.1 give the contribution of each food component to thermal diffusivity of
sweetpotato at a particular temperature. Moisture content variation was
simulated by varying the percentage by volume of water in equation (3.7), which
76
already included the effect the effect of temperature variation from the
mechanistic models, using Excel.
Thermal diffusivity data was generated for temperature range of 20 to
180°C in steps of 10°C and moisture content range of 40 to 70% w.b. in steps of
1%. Thermal diffusivity of sweetpotato determined by this method generally
increased with increase in moisture content and temperature. Figure 4.6 shows
a 3D mesh plot of thermal diffusivity of sweet potato. The regression model
developed using stepwise selection of variables of SAS is shown in equation
(4.6). Predicted 1 range from the model for typical frying conditions (temperature
and moisture content) is 1×10-7 -1.3×10-7 m2/s. This was comparable with 1.1 ×
10-7 m2s-1 reported for sweetpotato by Wadsworth and Spadaro (1969).
Matthews and Hall (1968) reported 14
value of 1.33×10-7 m2/s and 1.37×10-7 m2/s
for raw and cooked potato respectively.
14
= 7.32×10-8 + 3.41×10-10T + 2.1×10-8m + 5.31×10-13T2
(4.6)
Where:
1 = thermal diffusivity (m2/s)
T = temperature (°C)
m = moisture content (w.b.)
An estimate of 1 was also made from thermal conductivity, specific heat
and density from equation (2.10) for comparison. The value compared well with
predicted vales from equation (4.6) above.
77
6.0e-8
8.0e-8
1.0e-7
1.2e-7
1.4e-7
1.6e-7
1.8e-7
1.6e-7
1.4e-7
Thermal diffu
2
sivity (m /s)
1.8e-7
1.2e-7
1.0e-7
8.0e-8
6.0e-8
160
Te 140
mp 120
er 100
atu
re 80
(°C
)
0.45
0.50
is)
as
b
0.60
et
t (w
n
0.65
e
nt
Co
0.70
e
r
istu
Mo
0.55
60
40
20
Figure 4.6: Thermal diffusivity of sweetpotato as a function of moisture content
and temperature (from mechanistic model).
4.5 Moisture content during frying
Moisture content during frying was measured at specific periods of 0, 30,
60, 120, 180, 240 and 300 s of frying. Separate experiments were run for whole
sample, crust part and the core part of the product. Three sample sizes and four
oil temperature were used in frying and the experiment was conducted in
duplicate. Table A3 and A4 show the result from the experiment. All results are
in decimal (wet basis). Initial moisture content in sweet potato were as high as
0.82. As expected, moisture content reduced during frying. The moisture loss
rate was quite high at the beginning of frying as moisture on the surface of the
sample evaporated. The moisture loss rate however reduced as frying
78
progressed. It was observed that rate of moisture loss was higher for a smaller
sample size and also higher for higher oil temperature (Figure 4.7 and 4.8). This
same trend was observed for the three groups of moisture content (whole, crust,
core) determined.
The rate of moisture loss during frying has been expressed as the ratio of
a driving force to resistance from the product (Rice and Gamble, 1989). The
driving force is provided by the conversion of water to steam by heat. As frying
temperature increased, the moisture content for the same frying time decreased
since an increase in temperature resulted in a higher kinetic energy for water
molecules leading to a more rapid moisture loss in form of vapour which
ultimately reduced the moisture content of the product (Farinu and Baik, 2005).
Also, heat conduction to the center of food is faster for a product with smaller
dimension (diameter to thickness ratio in this case) leading to water molecules
at the center of this smaller product having a higher kinetic energy at any
particular time during frying and therefore, experience higher moisture loss rate
than a bigger sample. Krokida and co-workers (2000) also demonstrated that oil
temperature has a negative effect on the moisture content of French fried
potatoes.
79
0.80
Moisture Content (w.b.)
2.5 cm
0.75
3.5 cm
4.0 cm
0.70
0.65
0.60
0.55
0.50
0
50
100
150
200
250
300
350
Time of frying (s)
Figure 4.7: Moisture content variation among sizes for sample (whole sample)
fried at 180°C.
Moisture content. (w.b.)
0.85
150°C
160°C
170°C
180°C
0.80
0.75
0.70
0.65
0.60
0
50
100
150
200
250
300
350
Time of frying (s)
Figure 4.8: Moisture content of big sample (D/L=4) during frying at four different
oil temperatures.
80
4.6 Temperature during frying
Temperature was measured at 5 different locations in the sample. Table
A5 and A6 show a summary of temperature measured during frying.
Temperature at the surfaces of the sample rose rapidly at the beginning of frying
and typically reached a stable value (usually about 6 to 9°C below the oil
temperature) after about 45 s of frying. The bottom surface temperature was
usually higher than the top surface temperature at the beginning of frying. The
top surface temperature however caught up with bottom counterpart after about
50 s of frying. This trend is believed to be due to the fact that the bottom surface
receives more direct heat than the top at the beginning of frying before
significant bubbling started. However, the occurrence of vapour bubbling close
to the surface of the sweetpotato after the initial period increased the heat
transfer coefficient and therefore the temperature at the top surface.
The trend of temperature profile at the center and the two intermediate
locations of the sample were quite similar. The temperature rose with progress
in frying until it reached the boiling point of water, 100°C. The temperature then
remained stable for some time since the heat supplied was being spent on
moisture evaporation at this stage. After some time, the temperature started to
increase again. Figure 4.9 and 4.10 shows the typical temperature curves during
frying. It is noteworthy to mention that the temperature of the intermediate
locations (between the surfaces and the center) remained at 100°C until the
sample center temperature has reached boiling point (100°C) and stayed there
for some time. This is understandable since the bulk of moisture lost from the
center will pass through these locations before it gets to the surface and is lost
81
in the oil as vapour. Temperature within the product was found to increase with
oil temperature no matter what two sizes are being compared. Average product
temperature during frying was also found to vary inversely with sample size. The
smallest sample had highest temperature for a particular time during frying and
vice versa. This is due to a smaller thermal gradient across the smaller sample
profile which means that heat is conducted faster within the product.
160
140
Temperature (°C )
120
100
80
top
t2
60
center
t4
40
bottom
20
0
0
50
100
150
200
Time of frying (s)
250
300
350
Figure 4.9: Temperature of medium sample (D/L=3.5) during frying at 150°C (D=
sample diameter, L= sample thickness). t2 and t4 are temperature of positions
along the sample thickness in between top and center, and bottom and center
respectively.
82
200
180
160
Temperature (°C)
140
120
100
80
top
t2
60
center
t4
40
bottom
20
0
0
50
100
150
200
250
300
350
Time of frying (s)
Figure 4.10: Temperature of medium sample (D/L=3.5) during frying at 180°C.
(D= sample diameter, L= sample thickness). t2 and t4 are temperature of
positions along the sample thickness in between top and center, and bottom and
center respectively.
4.7 Heat transfer coefficient
Average heat transfer coefficient, h was determined for each of the six
time periods (0-30, 30-60, 60-120, 120-180, 180-240 and 240-300 s) during
frying based on heat energy balance. Resulting h values are shown in Table A7
and A8. Heat transfer coefficient rose quickly at the beginning of frying and
peaked at about 50-80 seconds of frying depending on the oil temperature of
frying. There is significant bubbling due to rapid moisture loss at the early stage
of frying. The rate of increase of h is higher for a higher oil temperature. This is
83
due to the lower viscosity of oil at higher temperature and higher drying rate and
bubbling/oil agitation which further increases the heat transfer coefficient. The
maximum h reached during frying is higher for higher oil temperature (Figure
4.11). After reaching the maximum, h during frying decreases and stabilizes at
450-550 W/m2°C for the remaining period of frying.
Heat transfer coefficient at the latter period of frying is slightly higher for
frying done at lower oil temperature. This is probably due to some slightly
significant bubbling in this period for frying done at lower oil temperature since
the initial drying rate is lower. Generally, h is higher for smaller sample than for
bigger sample fried at the same temperature (Figure 4.12). This is due to higher
product volume temperature in smaller samples since dimension is smaller and
1000
2
Heat transfer coefficient (W/m °C )
heat transfer rate to product center is faster.
800
600
400
150°C
160°C
170°C
180°C
200
0
0
100
200
300
Time of frying (s)
Figure 4.11: Heat transfer coefficient determined for medium sample (D/L=3.5).
84
2
Heat transfer coefficient (W/m °C)
1000
800
600
400
2.5 cm
3.5 cm
200
4.5 cm
0
0
100
200
300
400
Time of frying (s)
Figure 4.12: Heat transfer coefficient determined for samples fried at 150°C.
4.8 Computer simulation of sweetpotato during frying
Heat transfer and moisture transfer during frying of sweetpotato was
simulated using the heat transfer and diffusion modules of COMSOL
multiphysics. Governing equations, boundary conditions and initial conditions
are as defined in chapter 3. Computer modeling was done through the graphical
user interphase and the MATLAB prompt of the software. The built-in equation
for diffusion in the software was given in equation (3.13) and was found to
suffice for the purpose of this study. However the heat transfer equation
(equation 3.11) and its associated boundary conditions were user-defined.
Figures 4.13 and 4.14 show typical domain plots for temperature and
moisture content respectively after 300 s of frying simulation.
85
Sample thickness (centre to surface) (m)
Sample radius (centre to surface) (m)
Figure 4.13: Simulated temperature profile of medium sample (D/L=3.5) fried at
Sample thickness (centre to surface) (m)
170°C for 300 s.
Sample radius (centre to surface) (m)
Figure 4.14: Simulated moisture content profile of medium sample (D/L=3.5)
fried at 170°C for 300 s.
86
4.8.1 Comparison between temperature profiles of experiment and
simulation
Two separate temperature profiles (sample surface and sample centre)
were used to fit simulation to the experiment. A perfect fit was obtained for the
surface temperature for all the cases. The centre temperature however did not
record the same total success. Centre temperature profile from the simulation
matched perfectly, in most cases, with that of the corresponding experiment until
it reached 100°C then there was deviation. While centre temperature from the
experiment climbed to 100°C and stabilized at 100°C for some time (60 -100 s)
before it started to rise again, centre temperature from the simulation observed a
shorter period of temperature stability before it started rising again. Some of the
reasons thought to be responsible for this difference are:
1. The assumption that thermophysical properties of the crust and core
regions are same during frying is not strictly true. The crust is a dry
porous matrix that serves as an insulating material during frying and its
thermophysical properties varied from that of the core. This affected the
simulation result. Simulating each section with different but correct
thermal properties would yield a better result.
2. During frying, there is condensation on the thermal probes inserted into
the sample as they have been at a lower temperature than the sample
being measured. The condensate insulated the probes and the
temperature measured might have been slightly understated.
Attempts were however made to strike the best balance possible in fitting
87
both data. Figures (4.15 and 4.16) shows comparison between experimental
and simulation temperature profiles for sample size 3.5 cm fried at 160°C.
Average deviation between experiment and simulation for surface temperature
is 2.49°C while average deviation for centre temperature is 9.5°C.
180
160
Temperature (°C)
140
120
100
80
60
40
Experimental data
Simulated data
20
0
0
50
100
150
200
250
300
350
Time of frying (s)
Figure 4.15: Comparison of sample’s surface temperature of experiment and
simulation for medium sample (D/L=3.5) fried at 160°C
88
140
120
Temperature (°C)
100
80
60
40
Experimental data
Simulated data
20
0
0
50
100
150
200
250
300
350
Time of frying (s)
Figure 4.16: Comparison of sample’s centre temperature of experiment and
simulation for medium sample (D/L=3.5) fried at 160°C
4.8.2 Comparison between experimental and simulated moisture content
Two separate moisture content profiles (sample surface - crust and whole
sample) were used to fit simulation to the experiment. Similar to the case of
temperature, a perfect fit was obtained for the crust moisture content for all the
cases. The whole sample moisture content’s fit however was not as perfect as
that of crust. Slight difference was observed at the early stage of frying between
experimental and simulation moisture content data, and this difference only
increased with frying time. Some of the reasons thought to be responsible for
this difference are:
1. Frying is a moving boundary problem in which a water evaporation front
is established on the product surface at the beginning of frying as water
89
leaves the sample and this front gradually approaches the sample centre
as frying progresses leaving behind a porous matrix called crust. This
means that there was evaporation taking place inside the sample as
frying progressed. Moisture vapour flux was assumed to diffuse from the
sample into the oil at the sample surface. Coupling vapour flux at the
sample surface as was done in this study might therefore not represent
the process perfectly as it does not totally cater for this gap in moisture
loss inside the sample as frying progresses. Introducing this factor into
the surface boundary condition would lead to a better result.
2. The difference in thermophysical properties of crust and core regions also
would affect the moisture loss rate. Determining the actual properties of
the crust region and simulating the process with these values would lead
to a better result.
Simulation was done in such a way as to make provision for these issues
and to obtain the best fitting possible from practical point of view. Figures (4.17
and 4.18) shows comparison between experimental and simulation moisture
content profiles for 3.5 cm sample fried at 150°C. Average deviation between
experiment and simulation for surface moisture content was 7.26 kg/m3 while
average deviation for whole sample was 28.34 kg/m3.
90
Moisture content (kg/m 3)
700
Experimental data
Simulated data
600
500
400
300
200
0
100
200
300
Time of frying (s)
Figure 4.17: Comparison of surface moisture content profiles of experiment and
simulation of medium sample (D/L=3.5) fried at 150°C.
700
3
Moisture content (kg/m )
600
500
400
300
200
Experimental data
Simulated data
100
0
0
50
100
150
200
250
300
350
Time of frying (s)
Figure 4.18: Comparison of core moisture content profiles of experiment and
simulation of medium sample (D/L=3.5) fried at 150°C
91
4.8.3 Mass transfer coefficient
Mass transfer coefficient values from the simulation are presented in
Table A9 and A10. The trend observed in the result is that hm increased at the
onset of frying as the sample gained heat and water from sample evaporates
starting with the surface moisture. The coefficient increased with time of frying
as the moisture loss rate increases. However towards the end of frying, hm
decreased as the moisture loss rate decreased. This is explained by the
decrease in moisture content of the sample at the latter stage of frying hence a
reduction in the vapour bubbling and oil agitation which had before that period
been driving the heat transfer coefficient and moisture loss rate. A typical hm
profile during frying is shown in Figure 4.19.
8.E-06
2
Mass transfer coefficient (kg/m /s)
7.E-06
6.E-06
5.E-06
4.E-06
3.E-06
150°C
160°C
170°C
180°C
2.E-06
1.E-06
0.E+00
0
50
100
150
200
250
300
350
Time of frying (s)
Figure 4.19: Mass transfer coefficient determined during simulation of medium
sample (D/L=3.5) frying.
92
4.9 Correlation between heat transfer coefficient, h and mass transfer
coefficient, hm
An attempt was made at developing a relationship between heat transfer
coefficient, h and mass transfer coefficient, hm. The trend observed for both
parameters were different. While h rose sharply at the commencement of frying
reaching a maximum before 100 s of frying and settling to an almost stable
value for the last 120 s of frying in most cases (Figure 4.11), hm rose gradually
and continued rising for the most part of frying and did not reach a maximum
until around 200 s of frying in most cases (Figure 4.19).
Heat transfer coefficient and mass transfer coefficient reached maximum values
at separate times during frying. Heat transfer coefficient, h and mass transfer
coefficient, hm data were analysed for possible relationship. No direct
relationship was found between them. Both parameters were found to be a
function of experimental factors (sample size, oil temperature and frying time) at
different levels and cannot be directly correlated in a general form. However, a
relationship was found to exist between maximum h and the corresponding hm at
that point. The relationship found between them is shown in equation (4.7) with
R2 of 0.91 and p value of <0.001. Summarised ANOVA table for the model is
shown in Table B1.
hm1 = −9 × 10 −11 hmax + 2 × 10 −7 hmax − 6 × 10 −5
2
(4.7)
Maximum h and maximum hm during frying were also correlated and the
relationship found to exist between them is shown in equation (4.8) with R2 of
0.83 and p value of <0.001. Summarised ANOVA table for the model is shown in
Table B2.
93
hm 2 = −8 × 10 −11 hmax + 1 × 10 −7 hmax − 5 × 10 −5
2
(4.8)
Where:
hmax = maximum heat transfer coefficient reached during frying (W/m2.°C)
hm1 = mass transfer coefficient at maximum heat transfer coefficient (kg/m2.s)
hm 2 = maximum mass transfer coefficient reached during frying (kg/m2.s)
Equations (4.7) can be used to determine the corresponding mass
transfer coefficient during sweetpotato frying from the value of the maximum
heat transfer coefficient. Equation (4.8) can be used to determine the maximum
mass transfer coefficient from the value of the maximum heat transfer coefficient
during sweetpotato frying. These equations can also be used to estimate a
rough approximation during deep fat frying of similar food product like potato.
4.10 Effects of experimental factors on h and hm
The effects of experimental factors (oil temperature, sample size and time of
frying) were investigated and SAS stepwise regression was used to select
significant variables in developing a regression model for h and hm as a function
of these factors. The models developed for h and hm are as shown in equations
(4.9) and (4.10) with R2 of 0.90 and 0.76 respectively. The p-values were <0.001
and 0.012 respectively. Summarised ANOVA tables for the models are shown
in Table B3 and B4.
h = 1.28Toil − 1.5∆T 2 + 0.36Toil ∆T − 1.2dToil − 10.5
(4.9)
hm = 9.3 × 10−8 Toil − 8.6 × 10−7 d − 1.9 × 10−7 ∆T + 2.1× 10−6 ∆T 2 − 6 × 10−6.
(4.10)
Where:
94
h = heat transfer coefficient (W/m2°C)
hm = mass transfer coefficient (kg/m2.s)
Toil = oil temperature (°C)
d = sample diameter to thickness ratio i.e. D/L
8T = Toil -Tsurface (°C), 8T is a physical parameter representing progression of
frying.
From experimental observations and developed models, h and hm were
found to have direct positive relationships with oil temperature and an inverse
relationship with sample size. Equations (4.9) and (4.10) can be used to predict
the heat transfer and mass transfer coefficients during deep fat frying of
sweetpotato or rough estimate of h and hm of products with similar properties if
the product temperature and moisture content are known. The applicability of
these equations will be within the moisture content and temperature range used
in developing them, 0.45-0.75 w.b. and 20-180°C respectively.
95
5. SUMMARY, CONCLUSIONS AND RECOMMENDATIONS
5.1 Summary and conclusions
The general objective of this thesis research was to study the relationship
between heat and mass transfer coefficients during deep fat frying of
sweetpotato. In doing this, the thermophysical properties of the crop had to be
determined. Thermal conductivity, k of sweetpotato increased during frying. As
frying progressed, the product lost moisture but temperature also increased. The
net effect is a rise in k from an average of 0.43 W/m°C at the inception of frying
to 0.54 W/m°C after about 30 s of frying and 0.5 W/m°C towards the end of
frying. This same low-high-low trend was observed for thermal diffusivity with a
range of 1×10-7 to 1.3×10-7 m2/s and specific heat with a range of 3145 - 3320
J/kg°C during frying.
Density determined from the conducted experiment was modeled as a
function of moisture content only. The range of sweetpotato density during frying
was 1093 to 1203 kg/m3. For sweetpotato sample before frying (moisture
content of 0.81 - 0.82 wet basis), density was close to that of water. During
frying, temperature at the surface of the sample rose sharply reaching a stable
value (6 - 9°C below oil temperature) within 30 s. Temperature inside the sample
rose gradually, usually stable for some time at 100°C as moisture evaporated
96
before further increase. Moisture loss rate was also rapid for the sample surface
than the core.
Heat transfer coefficient, h was estimated based on energy balance
during frying and was found to approach a maximum between 80 and 140 s of
frying depending on the oil temperature and sample size. The range of
maximum h reached was 700 - 850 W/m2°C. Maximum h reached varied directly
with oil temperature but inversely with sample size. h at latter period of frying
(200 - 300 s) was 450 - 550 W/m2°C. Mass transfer coefficient, hm determined
from finite element computer simulation was found to reach a maximum (4×10-6
to 7.2×10-6 kg/m2.s) after 200 s of frying in most cases. Both h and hm were
found to increase with increase in oil temperature but decrease with increase in
sample size. No general correlation exists between h and hm. However, a
positive quadratic correlation existed between maximum h and maximum hm
during frying. Maximum h also varies directly with the corresponding hm at that
point. All models and empirical correlations developed in this study as a function
of temperature and moisture content are valid for typical frying condition range
(Temperature: 20 - 180°C and Moisture content: 0.45 - 0.75 (w.b.))
5.2 Recommendations
The relationship between heat transfer coefficient, h and mass transfer
coefficient, hm during deep fat frying of sweetpotato was investigated in this
study. It is expected that the study would provide an insight into the general
pattern of correlation between h and hm during frying and most other drying
processes. Specific situation for other processes and food products, however,
97
had to be studied since processes differ and thermophysical properties vary
from food to food. Frying is a moving-boundary problem in which a previously
non-existent moisture evaporation front is established on the food surface and it
moves towards the centre during frying. This phenomenon accounts for the crust
and the core parts of the sample which were not separated in this study. Their
properties actually vary slightly and this can be studied and incorporated into the
simulation in a future study. This could be achieved by incorporating separate
governing equations for heat and mass transfer for the crust and core regions
which will take into consideration the difference in their thermophysical
properties.
There are usually structural changes during deep fat frying. This is in form
of puffing after about 30 seconds of frying and shrinkage of sample towards the
end of frying. The simulation work in this study did not account for structural
change beyond the change in density during frying. It is recommended that a
future study incorporates this effect. Finally, for optimization of the frying process,
more work is required on energy consumption model and kinetics of quality
changes in the product (color, texture, nutritional index, etc.) and the results
should be coupled with the transport phenomena model.
98
REFERENCES
AOAC. 2002. AOAC official method 984.25. In Official Methods of Analysis of
AOAC International, 17th ed., Vol. 2, Gaithersburg, MD: The Association of
Official Analytical Chemists International.
Arce, J.A., V.E. Sweat and K. Wallapapan. 1980. Thermal diffusivity and
conductivity of soy flour. ASAE Paper No. 6528. St. Joseph, MI. American
Society of Agricultural Engineers.
Ateba, P. and G.S. Mittal. 1994. Modelling the deep fat frying of beef meatballs.
International Journal of Food Science and Technology. 29: 429-440.
Baik, O.D. (2004). Finite Element Method. In ABE898: Modelling of Food
Process Lecture Notes. University of Saskatchewan.
Baik, O. D. and G.S. Mittal. 2002. Heat transfer coefficients during deep-fat
frying of tofu disc. Transactions of the American Society of Agricultural
Engineers. 45: 1493-1499.
Bengtsson, N.E. and B Jakobsson 1974. Microwave Energy Applications
Newsletter 6(3).
Brailsford, A.D. and K.G. Major. 1964. Thermal conductivities of aggregates of
several phases, including porous materials. British Journal of Applied Physics.
15: 313-319.
Brodkey, R.S. and H.C. Hershey. 1988. Transport Phenomena: A Unified
Approach, Nueva York: McGraw-Hill. 516-520.
Buhri, A.B. and R.P. Singh 1993. Measurement of food thermal conductivity
using differential scanning calorimetry. Journal Food Science. 58: 1145.
Califano, A.N. and A. Calvelo. 1991. Thermal conductivity of potato between 50
and 100°C. Journal of Food Science. 56(2):586-587
Charney, P. and R.A. Seelig. 1967. Fruit & Vegetable Facts & Pointers:
Sweetpotatoes. Washington, D.C.: United Fresh Fruit & Vegetable Association,
Chen, H., B.P. Marks and R.Y. Murphy. 1999. Modeling coupled heat and mass
transfer for convection cooking of chicken patties. Journal of Food Engineering.
42: 139-146.
99
Choi, Y. and M. R. Okos. 1983. The thermal properties of tomato juice
concentrates. Transactions of the ASAE, 26(1): 305-311.
Choi, Y and M.R. Okos. 1985. Effects of temperature and composition on the
thermal properties of foods. In Food Engineering and Process Applications. Vol.
1. ed. M. Lemaguer and P. Jelen, 93. London, UK: Elsevier.
COMSOL™. 2005. COMSOL™ Multiphysics: Model Library - Heat Transfer
Module. Los Angeles. 192-202.
Costa, R.M., F.A.R. Oliveira, O. Delaney and V. Gekas. 1999. Analysis of the
heat transfer coefficient during potato frying. Journal of Food Engineering, 39(3):
293-299.
Derdour, L. and H. Desmorieux. 2004. A model for internal moisture diffusivity
during the regular regime: Comparison with experimental data obtained on
plaster and spirulina. Proceedings of the 14th International Drying Symposium.
São Paulo, Brazil, 22-25 August, A: 718-725.
Drouzas, A.E., Z.B. Maroulis, V.T. Karathanos and G.D. Saravacos. 1991. Direct
and indirect determination of the effective thermal diffusivity of granular starch.
Journal of Food Engineering, 13(2): 91-101.
Farkas, B.E. and L.J. Hubbard. 2000. Analysis of heat transfer during immersion
frying. Drying Technology, 18(6): 1269-1285.
Farkas, B.E., R.P. Singh and T.R. Rumsey. 1994. Modeling the effect of
properties and process conditions in food frying. Annual meeting of Institute of
Food Technologist. Atlanta, GA.: Institute of Food Technologists.
Farkas, B.E., R.P. Singh and T.R Rumsey. 1996a. Modeling heat and mass
transfer in immersion frying: model development. Journal of Food Engineering,
29: 211–226.
Farkas, B.E., R.P. Singh and T.R. Rumsey. 1996b. Modeling heat and mass
transfer in immersion frying: model solution and verification. Journal of Food
Engineering, 29: 227–248.
Farinu, A. and O. –D. Baik. 2005. Deep fat frying of foods: Transport
phenomena. Food Reviews International. 21: 389-410.
Farral, A.W., A.C. Chen, P.Y. Wang, A.M. Phanak, T. Hendrick and D.R.
Heldman. 1970. Thermal conductivity of dried milk in a packed bed.
Transactions of ASAE, 13(3): 391-394.
100
Gamble, M.H. and P. Rice. 1987. Effect of pre-frying drying on oil uptake and
distribution in potato crisp manufacture. International Journal of Food Science
and Technology, 22: 535–548.
Gamble, M.H. and P. Rice. 1988. Effect of initial tuber solid content on final oil
content of potato chips. Lebensmittel Wissenschaft und Technologie, 22: 62-65.
Gupta, T.R. 1990. Specific heat of Indian unleavened flat bread (chaphati) at
various stages of cooking. Journal of Food Process Engineering, 13(2): 217-227.
Hubbard, L.J. and B.E. Farkas. 1999. A method of determining the convective
heat transfer coefficient during immersion frying. Journal of Food Process
Engineering, 22: 201-214.
Hwang, M.P. and K. Hayakawa. 1979. A specific heat calorimeter for foods.
Journal of Food Science. 44(2): 435-438.
Kazarian, E.A. and C.W. Hall. 1965. Thermal properties of grains. Transactions
of ASAE, 8(1): 33-48.
Kestin, J and W.A. Wakeham. 1988. Transport Properties of Fluids: Thermal
Conductivity, Viscosity and Diffusion Coefficient Vol. I-1. Cindas data series on
material properties. ed. C.Y Ho, 221. New York, NY: Hemisphere Publishing
Corporation.
Kramkowski, R., 2
!
2and M. Serowik. 2001. Effect of moisture and
temperature of garlic on its specific heat. Electronic Journal of Polish Agricultural
Universities: Agricultural Engineering Series, 4(2).
Krokida, M.K., V. Oreopoulou and Z.B. Maroulis. 2000. Water loss and oil
uptake as a function of frying time. Journal of Food Engineering. 44: 39-46.
Matthews, F.V. Jr. and C.W. Hall. 1968. Method of finite differences used to
relate changes in thermal and physical properties of potatoes. Transaction of
ASAE, 11: 558.
Miller, K.S., R.P. Singh and B.E. Farkas. 1994. Viscosity and heat transfer
coefficients for canola, corn, palm and soybean oil. Journal of Food Processing
and Preservation, 18: 461-472.
Mittelman, N., S. Mizrahi and Z. Berk. 1984. Heat and mass transfer in frying. In
Engineering and Food, Engineering Sciences in the Food Industry. Volume I. ed.
McKenna B., 109–116. London, UK: Elsevier Science ad Technology Books.
Mohsenin, N.N. 1970. Physical Properties of Plant and Animal Materials New
York, NY: Gordon and Breach Science Publishers.
101
Moriera. R. G., Barrufet, M.A. and M.E. Castell-Perez. 1999. Deep Fat Frying:
Fundamentals and Applications. New York, NY: Kluwer Academic Publishers.
Moreira, R.G., J. Palau, V.E. Sweat and X. Sun. (1995a). Thermal and physical
properties of tortilla chips as a function of frying time. Journal of Food
Processing and Preservation, 19: 175-189.
Moreira, R. G., J.E Palau and X Sun. 1995b. Deep fat frying of tortilla chips: an
engineering approach. Food Technology, 49: 146-150.
Murakami, E.G. and M.R. Okos. 1989. Measurement and predictions of thermal
properties of foods. In Food Properties and Computer Aided Engineering of
Food Processing Systems. ed. R.P. Singh and A. G. Medina. New York, NY:
Kluwer Academic Publishers.
Ni, H. and A.K. Datta. 1999. Moisture, oil and energy transport during deep-fat
frying of food materials. Transactions of Institution of Chemical Engineers. Part
C: Food and Bioproducts Processing, 77: 194–204.
Picha, D. H. 1985. Crude protein, minerals and total carbtenoids in sweet
potatoes. Journal of Food Science, 50: 1768.
Rahman, S. 1995. Food properties handbook. 192. New York, NY: CRC Press.
Rice, P. and M.H. Gamble. 1989. Technical note: modeling moisture loss during
potato slice frying. Journal of Food Science and Technology, 24: 183–187.
Rice, P., J.D. Selman and R.K. Abdul-Rezzak. 1988. Effect of temperature on
thermal properties of ‘Record’ potatoes. International Journal of Food Science
and Technology. 23:281-286.
Rodriguez, R., B.L. Raina, E.B. Pantatisco and M.B. Bhatti. 1975. Quality of
Raw Material for Processing in Post harvest Physiology: Handling and Utilization
of Tropical and Subtropical Fruits and Vegetable. ed. E.B. Pantatisco, 467.
Westport CT: AVI Publishing Co.
Sabapathy N. D. 2005. Heat and mass transfer during cooking of chickpeameasurements and computational simulation. Unpublished M.Sc Thesis.
Saskatoon, SK: University of Saskatchewan.
Saravacos, G.D. and M.N. Pilsworth. 1965. Thermal conductivity of freeze-dried
model food gels. Journal of Food Science, 30: 773-778.
Scheerlinck, N., B.M. Nicolai, P. Verboven and J. De Baerdemaeker. 1996.
Finite element analysis of coupled heat and mass transfer problems with
102
random field material properties. ASAE Paper No. 96-3028. St. Joseph, MI:
American Society of Agricultural Engineers.
Siebel, E. 1892. Specific heat of various products. Ice and Refrigeration. 2: 256257.
Stitt, F. and E. K. Kennedy. 1945. Specific heats of dehydrated vegetables and
egg powder. Food Research, 10: 426- 436.
Stroshine, R. 1998. Physical Properties of Agricultural Materials and Food
Products. West Lafayette, IN: Purdue University.
Sweat, V.E. 1986. Thermal properties of foods, In Engineering Properties of
Foods. ed. M.A Rao, and S.S.H. Rizvi, 49-87. New York, NY: Marcel and
Dekker Inc.
Tang, J., S. Sokhansanj, S. Yannacopoulos and S. O. Kasap. 1991. Specific
heat capacity of lentil seeds by differential scanning calorimetry. Transactions of
ASAE. 34(2): 517-522.
Tong, L.S., Y.S. Tang. 1997. Boiling Heat Transfer and Two-Phase Flow, 2nd
edition. Washington D.C.: Taylor and Francis.
USDA. 1963. Agriculture Handbook No. 8: Composition of Foods; Raw,
Processed, Prepared. Washington, DC. United States Department of Agriculture.
USDA. 1984. Potatoes and Sweetpotatoes: Final Estimate. Statistical Bulletin
1978-1982. 709. Washington, DC.
Wadsworth, J.I. and J.J. Spadaro. 1969. Transient temperature distribution in
whole sweet potato roots during immersion heating. Food Technology, 23: 219.
Wang, N. and J.G. Brennan. 1992. Thermal conductivity of potato as a function
of moisture content. Journal of Food Engineering, 17: 153-160.
Wang, N. and J.G. Brennan. 1993. The influence of moisture content and
temperature on the specific heat of potato measured by differential scanning
calorimetry. Journal of Food Engineering, 19: 303-310.
Woodside, W. and J.H. Messmer. 1961. Thermal conductivity of porous media. I.
Unconsolidated sands. Journal of Applied Physics. 1688.
103
APPENDICES
APPENDIX A
Table A1. Density of sweetpotato determined by Multipycnometer.
I
II
1126
1189
1241
1298
1122
1171
1218
1312
Moisture content (w.b)
0.75
0.65
0.55
0.45
Trials
III
IV
3
Density (kg/m )
1125
1182
1263
1284
V
1118
1194
1233
1301
1124
1201
1255
1277
C.V = 1.03, RMSE = 12.49
Table A2. Specific heat of sweetpotato determined by DSC.
M.C
0.75
Trials I
II
0.65
I
II
0.55
I
II
0.45
I
II
Temp. (°C)
20
30
40
50
60
70
80
90
100
110
120
130
140
150
160
170
180
3088.1
3123.5
3164
3259.7
3292.1
3303.4
3313.8
3323
3329.7
3350
3385.2
3392.5
3399.5
3429.1
3477
3523.8
3566.2
2912.4
2933.4
2968.4
3119.9
3158.4
3124.2
3165.0
3165.9
3190
3261.8
3283.7
3301.4
3378.4
3416
3481.2
3498.3
3539.4
2864.5
2889.7
2902.4
2957.2
3003.5
3020.5
3035.8
3056.6
3100.4
3167.8
3220.8
3221.8
3240
3301.8
3342.5
3377.3
3401.5
2881.6
2901.4
2924.5
2979.8
2944.2
2925.3
2970.8
2975.6
2993.5
3090.6
3143.2
3170.8
3216.4
3275.4
3333.5
3381.5
3397
C.V = 4.33, RMSE = 123
104
2458.2
2486.7
2503.1
2581.1
2623.9
2638.4
2644.3
2657.2
2804
2870.6
2940.5
2933.5
2970.6
3016.1
3051.6
3122.4
3201.1
2424.6
2434.5
2454.6
2485.5
2557.4
2599.5
2631.3
2660.4
2816.3
2910.5
2963.2
2973.4
2975.4
3024.8
3081.4
3111.6
3176.9
2297.8
2314.6
2339.2
2340.2
2364.1
2409.3
2454.4
2503
2626.4
2754.1
2881.3
2891
2900.1
2954.6
3002
3044.5
3089.4
2285
2294.6
2308.4
2351.8
2403.1
2450.8
2472
2482.3
2630.7
2791.5
2840.7
2867.8
2888
2966.6
2998.4
3065.1
3077.6
Table A3. Moisture content of sweetpotato during frying (decimal, wet basis) –
Trial1
Temp.
(°C)
Sizes
Whole
Time (s)
150
160
170
180
I
II
III
I
II
III
I
II
III
I
II
III
30
60
120
180
240
300
Crust
Time (s)
0.791
0.760
0.729
0.697
0.675
0.667
0.780
0.748
0.721
0.689
0.664
0.647
0.781
0.748
0.721
0.687
0.657
0.649
0.786
0.758
0.726
0.694
0.670
0.656
0.782
0.746
0.717
0.679
0.648
0.637
0.784
0.746
0.717
0.678
0.647
0.641
0.790
0.760
0.709
0.681
0.658
0.634
0.778
0.741
0.704
0.661
0.641
0.633
0.780
0.744
0.702
0.667
0.641
0.633
0.749
0.705
0.658
0.611
0.587
0.583
0.764
0.722
0.683
0.646
0.620
0.614
0.777
0.732
0.691
0.654
0.631
0.626
30
60
120
180
240
300
Core
Time (s)
0.676
0.601
0.537
0.423
0.338
0.327
0.748
0.668
0.577
0.466
0.387
0.343
0.760
0.689
0.586
0.455
0.376
0.363
0.689
0.595
0.474
0.358
0.297
0.266
0.721
0.654
0.517
0.412
0.333
0.272
0.701
0.680
0.557
0.406
0.358
0.312
0.693
0.532
0.382
0.271
0.244
0.233
0.698
0.614
0.468
0.335
0.279
0.256
0.702
0.616
0.473
0.332
0.314
0.304
0.631
0.425
0.278
0.234
0.228
0.208
0.666
0.505
0.381
0.279
0.253
0.234
0.682
0.531
0.394
0.308
0.283
0.243
0.818 0.815 0.813 0.819 0.806 0.813
0.803 0.811 0.810 0.818 0.799 0.805
0.799 0.795 0.805 0.802 0.798 0.799
0.798 0.789 0.798 0.795 0.787 0.796
0.781 0.787 0.783 0.771 0.783 0.791
0.779 0.780 0.781 0.769 0.779 0.787
Sizes: I is (D/L=2.5), II is (D/L=3.5), III is (D/L=4)
0.803
0.799
0.798
0.782
0.776
0.772
0.803
0.799
0.796
0.788
0.783
0.779
0.813
0.804
0.793
0.788
0.782
0.773
0.802
0.798
0.786
0.779
0.769
0.762
0.801
0.799
0.791
0.786
0.776
0.762
0.815
0.799
0.794
0.780
0.777
0.762
30
60
120
180
240
300
105
Table A4. Moisture content of sweetpotato during frying (decimal, wet basis) –
Trial 2
Temp.
(°C)
Sizes
Whole
Time (s)
150
170
180
II
III
I
II
III
I
II
III
I
II
III
30
0.783
60
0.744
120
0.708
180
0.682
240
0.662
300
0.651
Crust
Time (s)
0.775
0.735
0.700
0.671
0.648
0.643
0.794
0.759
0.723
0.696
0.670
0.643
0.781
0.751
0.723
0.698
0.670
0.645
0.779
0.737
0.706
0.678
0.651
0.641
0.791
0.754
0.716
0.682
0.652
0.646
0.786
0.760
0.716
0.684
0.657
0.633
0.782
0.752
0.709
0.662
0.636
0.631
0.789
0.751
0.713
0.677
0.648
0.642
0.754
0.709
0.661
0.616
0.591
0.589
0.768
0.721
0.687
0.647
0.621
0.615
0.781
0.738
0.693
0.663
0.636
0.629
30
0.731
60
0.626
120
0.521
180
0.422
240
0.341
300
0.320
Core
Time (s)
0.750
0.652
0.579
0.470
0.391
0.347
0.764
0.694
0.612
0.459
0.380
0.367
0.693
0.568
0.478
0.361
0.285
0.269
0.724
0.641
0.535
0.417
0.335
0.35
0.724
0.683
0.559
0.410
0.362
0.337
0.697
0.536
0.386
0.327
0.276
0.237
0.702
0.619
0.470
0.368
0.282
0.261
0.706
0.621
0.478
0.350
0.312
0.286
0.636
0.489
0.324
0.239
0.233
0.225
0.670
0.509
0.381
0.282
0.255
0.238
0.686
0.553
0.425
0.312
0.297
0.247
0.807 0.810
0.813 0.804 0.806 0.811
0.803 0.809
0.809 0.802 0.805 0.805
0.801 0.7800 0.803 0.792 0.798 0.799
0.796 0.793
0.797 0.782 0.789 0.793
0.785 0.786
0.788 0.774 0.786 0.791
0.784 0.779
0.785 0.773 0.778 0.781
Sizes: I is (D/L=2.5), II is (D/L=3.5), III is (D/L=4)
0.801
0.796
0.787
0.780
0.772
0.771
0.808
0.799
0.796
0.787
0.784
0.773
0.812
0.803
0.794
0.790
0.781
0.773
0.800
0.796
0.784
0.777
0.770
0.766
0.799
0.794
0.790
0.777
0.766
0.762
0.804
0.798
0.795
0.785
0.771
0.766
30
60
120
180
240
300
I
160
106
Table A5. Temperature of sweetpotato sample during frying - Trial 1
Temp.
(°C)
Sizes
Top
Time (s)
150
I
180
II
III
I
II
III
I
II
III
0
23.05 22.53 23.71
30
125.3 126.4 119.0
60
134.2 136.3 130.2
120
139.8 140.1 134.2
180
142.6 143.0 135.6
240
143.8 141.9 135.7
300
143.7 142.7 136.2
T2 (Between Top and Center)
Time (s)
21.58
135.2
143.9
149.5
149.9
147.5
148.3
22.73
149.5
152.1
152.2
151.9
152.8
153.2
23.01
145.3
149.5
152.6
151.1
150.2
150.2
22.91
151.2
160.5
160.4
161.4
161.5
163.0
23.04
150.5
157.0
160.0
161.3
161.5
161.9
21.60
157.3
159.8
163.0
162.3
162.2
163.6
22.60
161.3
165.6
167.4
169.3
170.6
169.5
22.81
158.5
165.3
166.2
168.0
170.0
170.6
22.78
158.6
163.3
168.3
167.8
167.9
170.2
0
30
60
120
180
240
300
Center
Time (s)
22.19
71.13
93.75
103.4
112.1
125.8
133.5
22.03
71.52
89.00
100.8
114.6
125.9
132.8
22.21
78.05
95.35
101.0
106.4
115.8
124.9
22.35
69.60
94.75
103.0
112.6
126.0
136.5
22.82
75.12
91.52
105.1
113.6
124.4
131.4
21.03
76.25
93.30
101.8
116.8
130.6
136.5
22.60
71.60
94.90
104.8.
115.3
127.6
135.8
23.22
92.00
100.1
107.4
119.1
132.2
139.8
23.02
85.00
98.25
105.2
120.1
135.6
145.6
0
22.46 22.51 24.05 20.58
30
50.51 48.75 54.05 40.15
60
82.60 72.45 78.75 71.07
120
99.51 98.50 99.25 98.91
180
101.1 101.3 100.9 101.9
240
102.0 103.3 101.2 101.6
300
102.8 105.0 101.6 101.6
T4 (Between Center and Bottom)
Time (s)
23.19
42.79
69.46
98.30
101.4
101.6
102.6
22.99
47.11
73.35
99.61
100.9
101.5
103.2
22.05
53.83
80.75
101.7
103.8
104.9
106.4
23.58
63.12
85.75
100.5
102.7
103.9
105.2
21.11
36.07
67.80
98.40
101.7
102.9
106.0
22.35
49.69
81.25
103.1
103.2
104.9
107.3
22.90
62.34
85.00
101.0
102.5
104.7
108.8
22.71
53.30
80.50
102.6
104.8
108.1
113.0
0
30
60
120
180
240
300
Bottom
Time (s)
22.50
76.35
93.00
101.8
113.5
123.3
128.0
22.55
73.05
88.80
100.4
111.6
123.1
129.5
23.52
74.20
99.80
99.61
105.2
115.7
123.6
21.22
67.37
88.65
100.5
113.1
125.8
134.8
23.29
83.05
94.30
103.0
114.7
125.0
130.6
23.06
77.80
91.85
102.7
110.9
117.4
124.5
22.61
74.00
92.15
102.5
116.2
130.1
138.5
23.15
77.20
90.65
102.0
110.1
122.0
128.7
20.85
72.77
89.15
100.2
117.1
130.1
133.5
22.89
83.00
97.20
105.4
116.9
129.5
139.6
23.09
85.25
96.50
105.9
114.6
128.9
136.8
23.09
81.20
95.05
103.6
118.9
131.8
140.7
0
30
60
120
180
240
300
23.24
130.3
134.0
138.9
142.8
143.5
143.5
23.05
132.8
136.0
140.5
143.6
143.3
143.8
23.74
132.1
134.7
135.5
139.2
139.4
138.4
21.55
127.3
135.5
147.0
150.8
150.7
150.2
23.03
150.3
153.4
154.1
154.2
153.2
154.4
23.03
154.0
154.2
153.6
153.6
15.44
154.6
23.08
143.0
152.2
157.0
158.7
160.7
161.6
23.20
156.7
161.8
161.6
163.0
162.9
164.4
22.01
154.3
161.7
163.7
164.9
164.0
164.5
23.62
151.0
160.2
162.6
166.1
168.6
169.0
23.05
162.2
168.3
167.5
171.4
173.1
172.9
23.01
153.7
160.2
169.5
170.2
169.4
173.5
23.11
70.03
90.45
99.55
108.6
122.7
128.2
III
170
I
22.60
63.15
90.60
102.4
113.4
122.6
128.3
II
160
24.04
74.70
92.30
101.5
104.8
112.1
121.9
107
Table A6. Temperature of sweetpotato sample during frying - Trial 2
Temp.
(°C)
Sizes
Top
Time (s)
150
I
180
II
III
I
II
III
I
II
III
0
24.45 23.93 25.01
126.5 127.8 120.5
30
135.3 137.3 131.2
60
139.6 140.0 134.1
120
141.6 142.0 134.6
180
142.6 141.7 134.8
240
141.9 141.3 135.1
300
T2 (Between Top and Center)
Time (s)
22.98
136.2
144.8
149.4
149.2
147.2
147.8
23.60
149.1
151.6
151.4
151.1
152.8
152.1
23.21
145.7
150.6
153.4
151.2
150.8
150.7
23.91
151.2
161.1
161.5
161.9
162.1
162.3
23.04
151.2
156.9
160.8
160.9
160.9
161.8
22.58
158.3
159.8
162.5
161.9
162.7
162.9
23.75
161.4
165.5
167.6
169.5
170.1
169.9
23.05
158.9
165.4
165.9
167.8
169.6
169.8
23.07
158.5
163.2
167.4
168.2
168.3
169.5
0
30
60
120
180
240
300
Center
Time (s)
22.69
71.73
96.25
104.9
112.2
127.8
133.9
21.73
70.98
90.66
101.9
115.6
123.5
131.8
23.02
78.25
96.35
103.4
106.3
114.4
125.9
23.85
70.60
93.64
101.4
114.9
126.8
135.2
23.02
77.72
93.62
103.5
112.8
124.6
130.9
21.13
76.25
93.30
101.9
114.9
131.8
134.6
23.70
71.60
94.57
104.4
113.9
128.6
135.4
23.22
92.90
101.6
109.0
119.6
130.6
138.4
23.92
84.80
99.75
106.2
120.2
137.5
143.8
0
23.86 24.21 23.75 22.28
30
51.70 49.65 54.55 41.75
60
83.40 73.05 79.25 72.37
120
98.91 99.10 99.75 99.40
180
101.3 101.1 101.4 102.1
101.7 103.2 101.6 102.0
240
102.3 104.7 101.9 101.9
300
T4 (Between Center and Bottom)
Time (s)
23.59
43.49
69.86
98.88
100.9
101.5
102.1
23.79
47.91
74.15
101.1
101.4
102.0
102.5
22.85
55.33
81.35
103.1
103.5
104.6
105.9
24.58
64.12
87.25
101.1
102.5
103.2
104.8
21.71
35.99
69.33
99.32
101.8
102.3
105.6
22.95
50.49
81.85
104.5
104.7
105.3
106.6
22.56
62.96
85.66
101.7
103.3
105.1
107.1
23.62
53.80
81.60
102.9
104.6
107.8
113.4
0
30
60
120
180
240
300
Bottom
Time (s)
22.85
76.75
94.50
101.8
113.6
123.1
127.5
23.65
73.65
89.30
101.3
111.4
122.5
129.3
24.52
74.70
90.60
101.5
105.6
115.2
122.9
22.21
67.54
89.55
101.2
112.4
125.2
133.3
22.99
83.65
95.20
104.1
114.8
125.5
129.8
23.12
78.63
92.77
103.5
111.2
117.1
124.0
23.81
74.40
92.11
103.3
116.5
130.4
138.7
23.45
77.60
91.45
102.3
110.9
121.1
128.6
21.70
74.17
90.55
101.6
116.1
130.1
132.9
23.79
83.20
97.90
104.0
117.4
128.9
139.1
23.89
85.95
97.04
106.6
114.0
128.3
135.9
23.81
81.80
94.94
104.3
118.4
131.5
141.1
0
30
60
120
180
240
300
23.74
131.2
134.8
140.2
142.8
142.9
143.3
23.75
131.4
136.5
140.5
143.9
144.2
143.5
24.64
133.3
135.7
135.9
138.4
139.6
140.7
22.05
127.6
136.3
147.8
149.6
149.9
150.1
22.83
151.2
154.3
154.7
153.8
153.6
153.9
23.53
154.4
154.4
154.5
153.8
154.9
154.4
23.28
143.5
153.1
157.1
158.2
159.9
161.0
23.60
157.4
162.5
162.2
162.6
162.6
163.4
22.71
155.1
162.7
164.4
164.6
163.9
163.9
24.82
150.5
160.4
162.9
165.7
168.2
168.6
23.25
163.0
168.5
168.2
171.1
172.4
172.6
23.81
154.3
161.2
170.0
169.9
171.1
172.8
23.70
70.68
89.95
98.66
107.9
124.2
127.5
III
170
I
22.81
63.05
90.81
103.8
111.8
123.9
127.2
II
160
24.04
75.10
91.19
101.4
102.4
113.3
121.9
108
Table A7. Average heat transfer coefficient during frying (W/m2°C) - Trial 1
Temp.
150
160
(°C)
Sizes
I
II
III
I
II
III
Time (s)
30
136.8 103.0 103.0 146.9 109.5 109.3
60
597.3 599.8 583.5 644.0 596.3 597.9
120
696.1 680.5 714.0 722.9 713.1 726.0
180
551.2 531.6 585.9 546.7 488.4 493.0
240
474.3 481.9 562.2 518.5 483.1 502.1
300
441.1 467.5 538.9 502.1 439.0 486.0
Sizes: I is (D/L=2.5), II is (D/L=3.5), III is (D/L=4)
170
180
I
II
III
I
II
III
144.8
726.0
680.0
624.2
572.8
498.7
109.8
725.1
690.6
507.7
504.6
461.2
108.5
746.0
767.0
618.5
527.3
511.2
125.3
796.1
646.8
503.7
491.2
429.0
125.8
796.2
738.3
525.2
513.2
463.7
125.3
800.3
755.5
559.1
509.3
477.9
Table A8. Average heat transfer coefficient during frying (W/m2°C) - Trial 2
Temp.
150
160
(°C)
Sizes
I
II
III
I
II
III
Time (s)
30
131.2 100.3 102.3 139.5 107.7 108.3
60
615.8 585.9 489.7 688.6 593.5 589.5
120
710.7 700.6 674.5 715.2 697.1 720.2
180
512.9 503.5 527.5 476.5 545.8 551.7
240
508.0 496.3 539.6 554.1 459.5 447.1
300
400.3 451.8 512.5 505.5 473.7 436.8
Sizes: I is (D/L=2.5), II is (D/L=3.5), III is (D/L=4)
109
170
180
I
II
III
I
II
III
142.2
760.7
666.2
574.6
561.2
514.9
109.4
749.9
692.2
531.4
529.8
507.7
106.4
750.1
627.4
550.5
533.8
443.9
124.5
837.5
674.8
503.6
478.9
430.7
120.9
805.7
764.2
577.7
574.7
523.9
124.3
804.5
781.7
599.0
512.0
485.5
Table A9. Average mass transfer coefficient during frying (kg/m2.s) – Trial 1
Size
I
II
III
Temp. (°C)
150
Time (s)
0-30
30-60
60-120
120-180
180-240
240-360
2.70 × 10
-6
3.00 × 10
-6
3.00 × 10
-6
4.40 × 10
-6
4.90 × 10
-6
4.00 × 10
-6
1.40 × 10
-6
2.00 × 10
-6
2.20 × 10
-6
4.05 × 10
-6
4.00 × 10
-6
4.00 × 10
-6
9.00 × 10
-6
1.80 × 10
-6
2.20 × 10
-6
4.60 × 10
-6
4.50 × 10
-6
3.20 × 10
160
Time (s)
0-30
30-60
60-120
120-180
180-240
240-360
2.70 × 10
-6
3.20 × 10
-6
4.50 × 10
-6
5.70 × 10
-6
5.40 × 10
-6
5.40 × 10
-6
1.00 × 10
-6
3.00 × 10
-6
3.00 × 10
-6
4.80 × 10
-6
4.70 × 10
-6
4.20 × 10
-6
2.00 × 10
-6
2.00 × 10
-6
2.60 × 10
-6
5.20 × 10
-6
3.90 × 10
-6
3.50 × 10
170
Time (s)
0-30
30-60
60-120
120-180
180-240
240-300
2.70 × 10
-6
6.20 × 10
-6
6.70 × 10
-6
7.00 × 10
-6
6.50 × 10
-6
6.00 × 10
-6
2.50 × 10
-6
2.20 × 10
-6
5.60 × 10
-6
5.90× 10
-6
5.90 × 10
-6
5.70 × 10
-6
2.10 × 10
-6
2.40 × 10
-6
4.50 × 10
-6
6.10 × 10
-6
4.20× 10
-6
4.20 × 10
-6
2.60 × 10
-6
4.10 × 10
-6
5.70 × 10
-6
6.60 × 10
-6
5.50 × 10
-6
5.50 × 10
180
Time (s)
-6
0-30
4.50 × 10
-6
30-60
6.40 × 10
-6
60-120
6.90 × 10
-6
120-180
7.30 × 10
-6
180-240
6.70 × 10
-6
240-300
5.40 × 10
Sizes: I is (D/L=2.5), II is (D/L=3.5), III is (D/L=4)
110
3.90 × 10
-6
5.10 × 10
-6
5.50 × 10
-6
6.30 × 10
-6
7.20 × 10
-6
5.90 × 10
-6
-6
-6
-6
Table A10. Average mass transfer coefficient during frying (kg/m2.s) – Trial 2
Size
I
II
III
Temp. (°C)
150
Time (s)
0-30
30-60
60-120
120-180
180-240
240-360
2.00 × 10
-6
2.40 × 10
-6
3.30 × 10
-6
4.60 × 10
-6
5.30 × 10
-6
4.00 × 10
-6
1.40 × 10
-6
2.10 × 10
-6
2.20 × 10
-6
4.00 × 10
-6
4.00 × 10
-6
4.00 × 10
-6
9.00 × 10
-6
1.60 × 10
-6
1.80 × 10
-6
4.40 × 10
-6
4.40 × 10
-6
3.00 × 10
160
Time (s)
0-30
30-60
60-120
120-180
180-240
240-360
2.60 × 10
-6
4.10 × 10
-6
4.30 × 10
-6
5.80 × 10
-6
5.60 × 10
-6
5.50 × 10
-6
1.40 × 10
-6
3.00 × 10
-6
3.00 × 10
-6
4.70 × 10
-6
4.70 × 10
-6
4.20 × 10
-6
1.40 × 10
-6
1.50 × 10
-6
2.90 × 10
-6
5.20 × 10
-6
3.90 × 10
-6
3.50 × 10
170
Time (s)
0-30
30-60
60-120
120-180
180-240
240-300
2.70 × 10
-6
6.20 × 10
-6
6.50 × 10
-6
6.90 × 10
-6
6.40 × 10
-6
6.00 × 10
-6
2.30 × 10
-6
2.40 × 10
-6
5.00 × 10
-6
5.40 × 10
-6
5.70 × 10
-6
5.20 × 10
-6
1.80 × 10
-6
2.60 × 10
-6
4.50 × 10
-6
6.10 × 10
-6
5.00 × 10
-6
4.70 × 10
-6
2.60 × 10
-6
4.40 × 10
-6
5.50 × 10
-6
6.60 × 10
-6
5.40 × 10
-6
5.40 × 10
180
Time (s)
-6
0-30
4.50 × 10
-6
30-60
6.30 × 10
-6
60-120
6.70 × 10
-6
120-180
7.20 × 10
-6
180-240
6.70 × 10
-6
240-300
5.40 × 10
Sizes: I is (D/L=2.5), II is (D/L=3.5), III is (D/L=4)
111
3.90 × 10
-6
5.20 × 10
-6
6.20 × 10
-6
7.10 × 10
-6
6.10 × 10
-6
5.80 × 10
-6
-6
-6
-6
Table B1. Summarised ANOVA table for empirical correlation between heat
transfer coefficient, h and mass transfer coefficient, hm at maximum h.
Source
DF
h
h*h
1
1
Sum of Squares
37.7
5.4
Mean Square F Value Pr > F
37.7
5.4
720.3
102.6
<.0001
<.0001
D.F. of error = 21
Table B2. Summarised ANOVA table for empirical correlation between
maximum heat transfer coefficient, h and maximum coefficient, hm.
Source
DF
h
h*h
1
1
Sum of Squares
Mean Square F Value
45.5
8.7
45.5
8.7
599.2
88.3
Pr > F
<.0001
<.0001
D.F. of error = 21
Table B3. Summarised ANOVA table for effect of oil temperature, sample size
and frying progression on heat transfer coefficient during frying.
Source
DF
Toil
8T* Toil
8T2
d *Toil
1
1
1
1
Sum of Squares
Mean Square
F Value
Pr > F
55958.4
695955.8
1485064.4
15219.8
17.9
222.1
473.9
11.3
<.0001
<.0001
<.0001
0.0085
55958.4
695955.8
1485064.4
15219.8
D.F. of error = 115
Table B4. Summarised ANOVA table for effect of oil temperature, sample size
and frying progression on mass transfer coefficient during frying.
Source
Toil
d
8T
8T2
DF
1
1
1
1
Sum of Squares
Mean Square
1.77 × 10-10
6.34 × 10-11
1.57 × 10-10
4.98 × 10-12
1.77 × 10-10
6.34 × 10-11
1.57 × 10-10
4.98 × 10-12
D.F. of error = 115
112
F Value
210.81
75.27
86.10
5.92
Pr > F
<.0001
<.0001
<.0001
0.0163